10.05.2015 Views

Print Fluent Newsletter - MESco

Print Fluent Newsletter - MESco

Print Fluent Newsletter - MESco

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>Fluent</strong> News<br />

APPLIED COMPUTATIONAL FLUID DYNAMICS VOL XV ISSUE 1 • SPRING 2006<br />

ACADEMIC<br />

A Dry Passage<br />

to the Afterlife<br />

AUTOMOTIVE<br />

Citroën CS in<br />

a Crosswind<br />

Flexible<br />

Fliers<br />

ENVIRONMENTAL<br />

Activated Sludge<br />

Basins Get on Track<br />

FOOD<br />

Looking Inside<br />

Dough Mixers<br />

PROCESS INDUSTRIES<br />

SUPPLEMENT INSIDE!


EDITOR’S NOTE<br />

THE TERM “MULTIPHYSICS” has been a part of<br />

simulation engineers’ vocabulary for some time. It<br />

describes any situation where two or more physical<br />

phenomena are coupled together. Often, the term is<br />

used to describe simulations that involve both fluid<br />

and structural mechanics, such as fluid-structure interaction<br />

(FSI). During the past few years, we have run<br />

stories on FSI in <strong>Fluent</strong> News and in the current issue,<br />

several stories are featured that illustrate different<br />

approaches to this difficult engineering problem. The<br />

articles cover applications ranging from aerospace<br />

(p. 5) to healthcare (p. 10). Most describe tightly<br />

coupled interactions between CFD and structural<br />

solvers and one illustrates the use of MpCCI (Meshbased<br />

parallel Code Coupling Interface, from<br />

Fraunhofer SCAI) to manage the coupled calculation<br />

(p. 11). Not all such applications are tightly coupled<br />

however, and the Support Corner (p. 34) describes a<br />

tool, based on user-defined functions in FLUENT, that<br />

can be used to transfer data between fluid and structural<br />

solvers for loosely coupled interactions.<br />

In addition to the FSI stories, several of the articles in<br />

the Process Industries supplement could be given the<br />

multiphysics label as well. Multiphase flows and reacting<br />

flows are examples where strong coupling exists<br />

between either separate fluid phases (p. S13 - S14) or<br />

chemical species through temperature-dependent<br />

reactions (p. S8 - S14). Traditional processes such as<br />

mixing are covered (p. S3 - S6) as are novel processes<br />

that involve the flow through microchannels (p. S7) or<br />

the extrusion of foam products (p. S16).<br />

As is customary with <strong>Fluent</strong> News, the current slate of<br />

articles represents engineering efforts from many corners<br />

of the world. Air conditioning units being manufactured<br />

in Malaysia (p. 13) and aircraft fuselage design<br />

being performed in China (p. 19) are two examples.<br />

There are several articles from the US and Europe,<br />

covering topics such as automotive components (p. 16<br />

- 17), a unique dough conditioner (p. 22), a car in a<br />

crosswind (p. 14), and a two-stroke engine (p. 18). A<br />

few new applications of CFD are also presented. The<br />

flow inside an infant incubator is optimized (p. 24), and<br />

a ventilation system is installed in an Egyptian tomb to<br />

help preserve the ancient wall paintings (p. 28).<br />

Students shine once again with their unbound energy<br />

for innovation. This time, their efforts are used to develop<br />

a human-powered submarine (p. 29).<br />

Last November, a 64-bit version of FLUENT running on<br />

a Windows Cluster was showcased at Supercomputing<br />

2005. The performance of this exciting new capability<br />

is summarized (p. 30). In other product news, the<br />

upcoming releases of GAMBIT 2.3 and TGrid 4.0 are<br />

reviewed (p. 32). These new products and computing<br />

opportunities will make our work more manageable<br />

in the months and years to come. Please continue<br />

to keep us informed of your own efforts to push the<br />

limits of CFD and let us know what CFD has been able<br />

to do for you. <br />

LIZ MARSHALL<br />

fluentnews@fluent.com<br />

<strong>Fluent</strong> News is published by<br />

10 Cavendish Court • Lebanon, NH 03766 USA<br />

603 643 2600 • www.fluent.com<br />

© 2006 <strong>Fluent</strong> Inc. All rights reserved.<br />

ON THE COVER:<br />

Contours of presssure on the wing of the<br />

Aermacchi M346 Advanced Trainer<br />

Courtesy of Politecnico di Milano and Aermacchi SpA<br />

ON THE SUPPLEMENT COVER:<br />

Contours of velocity magnitude on a plane between<br />

two impellers in a stirred tank and vortex structures<br />

near the upper impeller, colored by velocity magnitude<br />

Courtesy of Prague Institute of Chemical Technology<br />

Editor: Liz Marshall<br />

Assistant Editor: Susan Wheeler<br />

Contributing Editors: Erik Ferguson and Keith Hanna<br />

Design: Lufkin Graphic Designs<br />

FLUENT, FiDAP, GAMBIT, POLYFLOW, G/Turbo, MixSim,<br />

FlowLab, Icepak, Airpak, and FloWizard are trademarks of<br />

<strong>Fluent</strong> Inc. Icepak and Airpak are joint developments of<br />

<strong>Fluent</strong> Inc. and ICEM-CFD Engineering. All other products<br />

or name brands are trademarks of their respective holders.


CONTENTS<br />

16<br />

20 13<br />

17<br />

8<br />

29 S3<br />

5<br />

S5 10<br />

31<br />

FEATURES<br />

APPLICATIONS<br />

5<br />

FLUID-STRUCTURE<br />

INTERACTION<br />

Flexible Fliers in the<br />

Transonic Regime<br />

13 HVAC<br />

8<br />

Unconventional Sail Design<br />

10<br />

Artificial Heart Valve Takes<br />

Shape<br />

16<br />

11 FSI Controls Flow Rate<br />

Fan Research Makes<br />

Cool Air Conditioners<br />

14 AUTOMOTIVE<br />

Citroën C5 in a<br />

Crosswind<br />

17<br />

Torque Converters<br />

Get In Gear<br />

Emissions Control<br />

Through Carbon<br />

Canisters<br />

18<br />

POWER TOOLS<br />

Scavenging in a<br />

Stratified, Charged<br />

Two-Stroke Engine<br />

19 AEROSPACE<br />

Nacelle Impact on<br />

Aircraft Wing &<br />

Fuselage<br />

20<br />

Compressors Benefit<br />

from the NASA<br />

Rotor 37<br />

21 FOOD<br />

Image-based Meshing:<br />

Easy as Pie<br />

22<br />

Looking Inside Dough<br />

Mixers<br />

24 HEALTHCARE<br />

CFD Assists Neonatal<br />

Intensive Care<br />

26 ENVIRONMENTAL<br />

Activated Sludge<br />

Basins Get on Track<br />

<strong>Fluent</strong> News · Spring 2006 3


CONTENTS<br />

DEPARTMENTS<br />

28<br />

ACADEMIC NEWS<br />

A Dry Passage to the Afterlife<br />

29<br />

Student Submariners Peddle<br />

Their Way to Victory<br />

14<br />

11<br />

29<br />

Convective Motions Inside a<br />

Gearbox<br />

30<br />

PRODUCT NEWS<br />

<strong>Fluent</strong> & Microsoft Team to<br />

Deliver 64-bit FLUENT on<br />

Windows Clusters<br />

31<br />

Quick Turnaround with<br />

Rapid Flow Modeling<br />

32<br />

Impressive Preprocessing<br />

33 PARTNERSHIPS<br />

Discrete Element Modeling<br />

of Particles for FLUENT<br />

34<br />

SUPPORT CORNER<br />

Mapping Thermal Data<br />

from FLUENT to Structural<br />

Codes Quickly<br />

23<br />

S7 32<br />

36<br />

AROUND FLUENT<br />

<strong>Fluent</strong> Opens Larger Office<br />

in Ann Arbor, MI<br />

17<br />

27<br />

36<br />

Upcoming User Group<br />

Meetings<br />

PROCESS INDUSTRY<br />

SUPPLEMENT<br />

S6<br />

s2 OVERVIEW<br />

Transporting CFD to<br />

Process Engineers<br />

s12 FURNACES<br />

Ultra-low NOx<br />

Burners Get Cracking<br />

S11<br />

S15<br />

s3 MIXING<br />

Tracing Homogenization<br />

s5<br />

Finding the Optimum<br />

Blend Time Calculation<br />

s7 MICROREACTORS<br />

Liquid Mixing<br />

in Microreactors<br />

s8<br />

THERMAL RUNAWAY<br />

Preventing Runaway<br />

Reaction Accidents<br />

s10 EMISSIONS<br />

Scrubbers for<br />

Flue Gas Cleanup<br />

s13 MULTIPHASE<br />

Controlling Droplet<br />

Size Distribution in<br />

Emulsions<br />

s14 Understanding<br />

Fluid-Bed Coating<br />

s15 PUMPS<br />

Pumping out New<br />

Designs More Quickly<br />

s16 EXTRUSION<br />

Polymer Processing<br />

Simulation for Foam<br />

Extrusion<br />

4 <strong>Fluent</strong> News · Spring 2006


FLUID-STRUCTURE INTERACTION<br />

Flexible Fliers<br />

in the Transonic Regime<br />

By L. Cavagna, G. Quaranta, P. Mantegazza, Department of Aerospace Engineering, Politecnico di Milano, Italy<br />

D. Marchetti, M. Martegani, M346 Program, Aermacchi SpA, Varese, Italy<br />

Pressure contours on the surface of the<br />

M346 Advanced Trainer<br />

FOR AN AIRCRAFT IN TRANSONIC FLIGHT, shock<br />

waves appear and move through the flow field. As a<br />

consequence of the unsteady flexible motion of the<br />

aircraft skin, shock waves dynamically modify the<br />

pressure distribution, eventually causing a drop in<br />

the flutter velocity, or speed at which the aircraft<br />

vibrations are no longer damped and instead grow<br />

in amplitude. The premature onset of flutter during<br />

transonic conditions, otherwise known as the “transonic<br />

dip” effect, is usually underpredicted by classical<br />

unsteady potential methods [1]. In a research<br />

project at the Politecnico di Milano [2], a computational<br />

aeroelastic (CA) model for deformable aircrafts<br />

in the transonic regime has been developed. Special<br />

care has been taken to create an environment that<br />

integrates all of the relevant physics, keeping in<br />

mind the high number of analyses that typically have<br />

to be run during the development of an aircraft. The<br />

implemented methods have been applied to the<br />

new generation Aermacchi M346 trainer.<br />

The aeroelastic solver core is implemented in FLUENT<br />

through the Scheme programming language.<br />

Access to flow variables and data manipulation is<br />

permitted through an extensive use of user-defined<br />

functions (UDFs) called at different times by the<br />

master Scheme library. The structural model comes<br />

from the commercial finite-element code MSC-<br />

Nastran, which has a long tradition in the aerospace<br />

industry.<br />

The adoption of a partitioned approach for the solution<br />

of fluid-structure interaction (FSI) problems<br />

requires the definition of an interface method to<br />

The Advanced Trainer M346 developed by Aermacchi<br />

<strong>Fluent</strong> News · Spring 2006 5


FLUID-STRUCTURE INTERACTION<br />

The aeroelastic procedure is illustrated for geometry elements (top) and the<br />

resulting mesh (bottom) for a case of amplified motion<br />

The AGARD 445.6 wing showing two modes of<br />

structural deformation<br />

exchange displacements and velocities from the<br />

structural grid to the aerodynamic wet surfaces of<br />

the CFD grid and to transfer back aerodynamic<br />

forces on the structural nodes. The structural and<br />

CFD models are described in a very different and<br />

often not compatible way; this is especially true in<br />

an industrial environment, where the models usually<br />

come from different departments. For the CA<br />

model developed at Politecnico di Milano, an interfacing<br />

procedure based on a “mesh-free” moving<br />

least squares (MLS) method [3] is used. This method<br />

is suitable for the treatment of geometrically complex<br />

configurations and, unlike many other interface<br />

methods, it ensures the conservation of momentum<br />

and energy transferred between the fluid and structure,<br />

a key factor for stability analysis like flutter. In<br />

fact, the comparison of spurious energy created or<br />

destroyed by the interface scheme may alter the<br />

stability boundary of the system.<br />

The primary assessment necessary for the aeroelastic<br />

certification of an aircraft is related to analyses of the<br />

linearized solutions to catch instabilities and specifically,<br />

flutter conditions. In the case of strong nonlinearities<br />

in the flow field, it is necessary to assess the<br />

stability of each movement associated with each<br />

equilibrium point of the aeroelastic system.<br />

Consequently, any flight configuration could potentially<br />

assume a different stability behavior. However, if<br />

there are no abrupt changes in the fluid flow, it is<br />

reasonable to consider the linearization around a specific<br />

flight condition as representative of the behavior<br />

at nearby flight points, which are characterized by<br />

small differences in the mass and stiffness distributions,<br />

and consequently small variations in the aircraft<br />

attitude. In the work presented by Melville [4] on an<br />

application similar to the one considered here, a<br />

stability investigation was conducted following a<br />

“numerical experiment” technique in which the<br />

stability is determined by analyzing the decaying or<br />

diverging behavior of time responses. This method is<br />

extremely time-consuming because a large number<br />

of analyses are required to bracket the flutter conditions.<br />

An approach based on the adoption of a CFD<br />

solver for generating reduced order models (ROMs) of<br />

transonic aerodynamics is more affordable and<br />

industrially oriented. As in classical flutter analyses,<br />

the result from this type of approach is a linearized<br />

model in either the frequency or time domain that<br />

can be used as an efficient tool for stability boundary<br />

assessment and dynamic loading response analysis.<br />

In order to avoid any possible source of error, a<br />

backup procedure should also be available to run<br />

the coupled nonlinear analysis. This alternative<br />

procedure can be used to verify and validate key<br />

instability points obtained by using the simplified<br />

linearized approach. A direct time integration can<br />

be easily implemented for this purpose using a<br />

loosely coupled, partitioned algorithm.<br />

6 <strong>Fluent</strong> News · Spring 2006


FLUID-STRUCTURE INTERACTION<br />

To correctly represent the structural deformation of<br />

the aircraft, the CFD computational grid must be<br />

modified at each time step in order to be compatible<br />

with the new aircraft shape. Grid deformation<br />

has a significant impact on the time required by<br />

CFD for aeroelastic simulations. For the sake of overall<br />

computational efficiency, nonlinear models for<br />

grid deformation have been avoided. Instead, an<br />

elastic analogy is used to represent the grid as a linear<br />

elastic continuum with a local Young’s modulus<br />

proportional to the minimal dimension of each element.<br />

The structural analogy provides ample freedom<br />

for choosing the material constitutive properties<br />

that will rule the grid deformation. Furthermore,<br />

by dividing the computational domain into different<br />

fluid zones, the deformation process can be customized<br />

by choosing different structural properties<br />

for the adjacent wall zones. The result is a flexible<br />

and extremely robust grid deformation algorithm.<br />

The structural model is represented as a set of<br />

modal shapes. To accelerate the solution time, the<br />

linearity of the problem is exploited, allowing the<br />

CFD grid deformation to be expressed as a superposition<br />

of grid deformations computed for each<br />

modal shape. All of the deformed grids associated<br />

with each modal shape are computed only once<br />

and stored in a database. Special care must be<br />

taken in the treatment of the governing surface<br />

movements since surface rotation can modify the<br />

CFD domain topology by creating cuts and new<br />

wall surfaces at the boundaries. As a result, the correct<br />

treatment of these conditions ideally requires<br />

the creation of a new mesh during the transient<br />

simulation, significantly increasing the computational<br />

burden. To efficiently overcome this hurdle, a<br />

solution based on a non-conformal mesh topology<br />

is used. This allows distinct cell domains to be<br />

defined that are associated with the movable surfaces<br />

that may be deformed independently from<br />

the rest of the grid. The implemented solver<br />

has been applied to the AGARD 445.6 wing, a<br />

classic aeroelastic benchmark case with published<br />

data [5]. The results are in good agreement with<br />

measurements in the transonic regime and correctly<br />

predict the transonic dip. <br />

References:<br />

1 Isogai, K.: On the Transonic-dip Mechanism of Flutter<br />

of a Sweptback Wing. AIAA Journal, 17:735-795, 1979.<br />

2 Cavagna, L., Quaranta, G., Ghiringhelli, G.L. and<br />

Mantegazza, P.: Efficient Application of CFD Aeroelastic<br />

Methods Using Commercial Software. In International<br />

Forum on Aeroelasticity and Structural Dynamics<br />

IFASD-2005, Munich, Germany, June 28 - July 1 2005.<br />

3 Quaranta, G., Masarati, P. and Mantegazza, P.: A<br />

Conservative Mesh-free Approach for Fluid-structure<br />

Interface Problems. Papadrakakis, M., Oñate, E. and<br />

Schrefler, B., Editors, in International Conference on<br />

Computational Methods for Coupled Problems in Science<br />

and Engineering, CIMNE, Santorini, Greece, 2005.<br />

4 Melville, R.: Nonlinear Mechanisms of Aeroelastic<br />

Instability for the F-16. AIAA Paper 2002-0871, January<br />

2002.<br />

5 Yates, E.C.: AGARD Standard Aeroelastic Configurations<br />

for Dynamic Response. I wing 445.6. R 765, AGARD,<br />

1985.<br />

V F<br />

Modal amplitude<br />

Modal amplitude<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

Experiment<br />

FLUENT<br />

0.2<br />

0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2<br />

Mach number<br />

The transonic dip, or decrease in flutter velocity index<br />

in the transonic regime, is nicely captured by the<br />

computational aeroelastic model<br />

.002<br />

.001<br />

.000<br />

-.001<br />

Mode 1<br />

Mode 2<br />

-.002<br />

0 8 16 24 32 40 48<br />

Structural dimensionless time<br />

.15<br />

.10<br />

.05<br />

.0<br />

-.05<br />

-.10<br />

Mode 1<br />

-.15<br />

Mode 2<br />

-.20<br />

0 4 8<br />

12 16<br />

Structural dimensionless time<br />

Modal amplitude histories for the AGARD 445.6<br />

wing for the case with Mach number 0.678 and<br />

flutter velocity index 0.34 (top) and 0.5 (bottom)<br />

Different fluid domains and a non-conformal mesh are used to control the motion of the various surfaces<br />

<strong>Fluent</strong> News · Spring 2006 7


FLUID-STRUCTURE INTERACTION<br />

Unconventional<br />

Sail Design<br />

By Paolo Conti, Marco Argento, Matteo Scarponi, Dipartimento Ingegneria Industriale, Università degli Studi di Perugia, Italy<br />

Tornado class catamaran<br />

Photo Courtesy of HoldenSailing<br />

IN THE WORLD OF DINGHY AND YACHT<br />

racing, the sail design process is largely<br />

based on the experienced eye of the sailmaker;<br />

with few exceptions, the refinement<br />

of a given sail shape is a time-consuming<br />

process and requires a great amount of fullscale<br />

tests at sea. One of the most relevant<br />

issues in sail design is the correlation of a<br />

nominal sail shape with its wind-loaded<br />

shape and its actual performance. The aim<br />

of the present work is to provide a partial<br />

response to such a problem through the<br />

development of a systematic procedure for<br />

analysis and comparison of different sail<br />

designs. In particular, the goals are to<br />

estimate how given design factors affect<br />

the sail’s performance and to provide an<br />

analytical model of the phenomenon. The<br />

method used is based on coupled structural<br />

and CFD simulations and the Design of<br />

Experiments (DOE) technique.<br />

According to Richter et al. [1], computational<br />

fluid dynamics “has demonstrated<br />

the ability to predict sail and appendage<br />

forces under upwind conditions.” In their<br />

work, the authors integrated FLUENT and<br />

a proprietary software for structural analysis<br />

to estimate the flying shape of sails.<br />

Thanks to CFD, an adequate degree of<br />

accuracy can also be obtained when predicting<br />

the behavior of a yacht sailing<br />

downwind [2].<br />

Regression techniques can be regarded as<br />

well-established tools in the naval architecture<br />

domain [3]. Furthermore, DOE has<br />

proved to be a valuable tool in sail design<br />

[4]. However, the potential of DOE techniques<br />

has not been fully explored. DOE<br />

offers three advantages [5]. First, a number<br />

of parameters (both discrete and continuous)<br />

and responses can be involved.<br />

Second, regression models can be provided<br />

with reasonably low computational effort.<br />

Third, possible design trends can be identified<br />

by means of appropriate representation<br />

of responses, or response surfaces, over a<br />

given design space.<br />

The selected case study is the design<br />

of asymmetrical spinnakers used by the<br />

Tornado, the Olympic class catamaran. A<br />

set of parameters representing a full 3D<br />

spinnaker shape has been identified and<br />

their domains evaluated, in order to perform<br />

systematic perturbations on a given<br />

sail design and set up a DOE-driven analysis.<br />

The surface mesh on the sail and wind<br />

tunnel boundaries<br />

A mesh sizing function whose source is<br />

the sail<br />

8 <strong>Fluent</strong> News · Spring 2006


FLUID-STRUCTURE INTERACTION<br />

Streamlines coloured by velocity magnitude; the<br />

spinnaker is viewed from the leeward side<br />

Comparison between a design shape (left) (input to the iterative process) and the corresponding flying shape<br />

(right) after convergence is reached following four fluid-structure iterations<br />

Although such parameters refer to a sail section at a<br />

given mast height, their perturbations have been<br />

propagated to the rest of the sail in order to obtain<br />

results consistent with common sailmaking practices.<br />

The selected process factors are the sail section<br />

camber, its position over the chord, and the<br />

section twist angle with respect to the sail foot<br />

chord (or bottom edge of the sail).<br />

Because of the need to sample a significant region<br />

of the design space, the various designs evaluated<br />

are considerably different from each other; notable<br />

differences have been observed in terms of the sail<br />

angle of attack at given reference heights. The<br />

structural behavior of the design shapes can’t be<br />

represented as that of rigid bodies. In fact, the<br />

overall sail shape undergoes large displacements<br />

and a fluid-structure interaction approach<br />

becomes necessary.<br />

A “numerical wind tunnel” consisting of a coupled,<br />

iterative procedure involving structural and fluiddynamic<br />

calculations has been developed to estimate<br />

the wind-loaded shapes of given sails and their<br />

performance in terms of driving force, side force,<br />

and heeling moment. At each iteration, FLUENT<br />

evaluates the pressure field over a given sail surface<br />

and provides it to a structural module that yields a<br />

new loaded shape. For the purposes of the structural<br />

analysis, the sail is modeled as an infinitely flexible<br />

web with fixed elongation [6]. Each structural<br />

step consists of a constrained optimization routine<br />

whose objective function is the work of aerodynamic<br />

forces acting on the sail.<br />

Since a wind tunnel facility suitable for the testing of<br />

model sails is available at the University of Perugia,<br />

the computational domain for the CFD analysis has<br />

been scaled accordingly, for validation purposes.<br />

The numerical wind tunnel underwent preliminary<br />

tests showing good agreement with actual flow field<br />

data. In order to be consistent with the actual<br />

process of sail trim, the ideal flow incidence with<br />

respect to a given reference section of the sail<br />

should be preserved until the iterative method converges.<br />

As an example, critical conditions may occur<br />

at low angles of attack, where the leading edge of<br />

the sail collapses easily.<br />

A proper module was therefore implemented<br />

through GAMBIT scripting language, which allowed<br />

the sail to be correctly positioned with respect to<br />

the flow field at the beginning of each FLUENT run,<br />

simulating actions taken by the helmsman. The<br />

problem just described implies that a unique meshing<br />

strategy must be suitable for several sail shapes<br />

(those generated at each step of the iterative<br />

process). The use of sizing functions allowed these<br />

difficulties to be overcome and, at the same time,<br />

provided a satisfactory mesh quality independent of<br />

the sail shape considered.<br />

Results of all simulations were collected and<br />

processed through an analysis of variance tools, in<br />

order to select the geometric parameters affecting<br />

sail performance. Regression models and response<br />

surfaces were then evaluated, in order to provide<br />

guidelines for future improvements of a given<br />

design. A validation of the present approach has<br />

been carried out by means of wind tunnel tests and<br />

consists of two stages. First, the performance in<br />

terms of driving force coefficient was evaluated.<br />

Second, the predicted and actual flying shapes of a<br />

given sail were compared. An optical measurement<br />

system [7] allowed the wind loaded shapes of<br />

model sails during wind tunnel runs to be digitized.<br />

In both cases, reasonable agreement with numerical<br />

results was found, which suggests the possibility<br />

of using the methodology described above as a<br />

valuable tool for sail designers. <br />

Response surface showing the influence of camber<br />

and position, while twist is fixed to its mid-range<br />

value, obtained from the final regression model<br />

References<br />

1 Richter H.J., Horrigan K.C. and Braun J.C.:<br />

Computational Fluid Dynamics for Downwind Sails. In<br />

Proc. of The 16th Chesapeake Sailing Yacht<br />

Symposium, 2003.<br />

2 Jones, P. and Korpus, R.: International America’s Cup<br />

Class Yacht Design Using Viscous Flow CFD. In Proc. of<br />

The 15th Chesapeake Sailing Yacht Symposium, pp.<br />

27-34, 2001.<br />

3 Battistin, D., Peri, D., and Campana, E.F.: Geometry<br />

and Resistance of the IACC Systematic Series Il Moro di<br />

Venezia. In Proc. of The 17th Chesapeake Sailing Yacht<br />

Symposium, pp. 33-51, 2005.<br />

4 Lasher, W.C., Sonnenmeier, J.R., Forsman, D.R., Zhang,<br />

C. and White, K.: Experimental Force Coefficients for a<br />

Parametric Series of Spinnakers. In Proc. of The 16th<br />

Chesapeake Sailing Yacht Symposium, 2003.<br />

5 Myers, R.H. and Montgomery, D.C.: Response Surface<br />

Methodology. John Wiley & Sons, New York, 2002.<br />

6 Le Maitre, O., Huberson, S.G. and Souza de Cursi, J.E.:<br />

Application of a Non-convex Model of Fabric<br />

Deformations to Sail Cut Analysis. J. Wind Engineering<br />

and Industrial Aerodynamics, Vol. 63, pp. 77-93, 1997.<br />

7 Barone, S., Bianconi, F., Conti, P., Razionale, A. and<br />

Scarponi, M.: Acquisition and Modelling of 3D Sail<br />

Shapes. CAD 2004, International CAD Conference and<br />

Exhibition, Pattaya Beach, Thailand, 2004.<br />

more.info@<br />

matteo.scarponi@unipg.it<br />

<strong>Fluent</strong> News · Spring 2006 9


FLUID-STRUCTURE INTERACTION<br />

t = 1.00 sec<br />

Artificial<br />

Heart Valve<br />

Takes Shape<br />

t = 1.04 sec<br />

t = 1.14 sec<br />

By Jan Vierendeels, Joris Degroote, Kris Dumont, Lieve Lanoye, Pascal Verdonck,<br />

Ghent University, Belgium<br />

AT GHENT UNIVERSITY, BELGIUM, research on fluid-structure interaction<br />

(FSI) modeling has been ongoing since 1996. At the beginning, in-house codes<br />

were developed to compute the interaction between the motion of the cardiac<br />

wall and the blood flow in the heart [1]. During the past ten years, CFD and CSD<br />

(computational structural dynamics) packages have made significant progress<br />

and nowadays, the best approach for coupling fluid motion and structural displacement<br />

is to use existing software packages as partitioned solvers. When using<br />

a partitioned approach, however, it can be difficult to obtain convergence, especially<br />

when the interaction between the different motions is strong. This is the<br />

case for blood flow in the heart and for blood flow through a heart valve with<br />

flexible leaflets. Both of these types of blood flow have been simulated using the<br />

FSI approach developed at Ghent University. The blood flow in the moving<br />

geometries is computed with FLUENT 6.2, and the structural problem is solved<br />

using in-house software. The coupled calculation is robust and does not cause<br />

undue convergence difficulty, even for strongly interacting system responses.<br />

During the calculation, an implicit coupling method is used. For each new time,<br />

a new position of the boundary and the corresponding load distribution on that<br />

boundary is sought. This information is the result of simultaneous solutions of the<br />

CFD and CSD solvers, and could be obtained by alternating the solver calls. For<br />

strongly coupled cases however, such an approach can only be stabilized by<br />

reducing the underrelaxation factors or by using an Aitken-like technique, both<br />

of which tend to slow the rate of convergence.<br />

t = 1.36 sec<br />

t = 2.00 sec<br />

The figures show a flexible leaflet being opened by<br />

an accelerating flow; the stiffness of the leaflet is very<br />

small, so that it undergoes a huge displacement from<br />

the closed to the fully open position and back again<br />

To speed the convergence, a reduced order model is used [2]. If the Jacobian of<br />

the fluid solver were known when solving the structural problem, one would<br />

know how the load would change when a certain boundary displacement is<br />

applied, so the load change would not need to be calculated with the fluid solver.<br />

The boundary condition for the structural solver (the load) would not need to be<br />

assumed constant during this call, but could be written as a function of the<br />

unknown boundary position itself, resulting in a much better prediction for the<br />

boundary position. Since the Jacobian of the fluid solver is not available, a<br />

reduced order model of the solver can be built for which a Jacobian can be<br />

constructed. This approximation of the Jacobian of the fluid solver is then used<br />

for the coupled calculation. During each time step, whenever FLUENT is called<br />

a boundary displacement mode is applied and the corresponding load change<br />

at the boundary is retrieved through user defined functions (UDFs). This<br />

information is used to build up the reduced order model for the fluid problem.<br />

The structural solver is then called with the variable load boundary condition.<br />

FLUENT is called again and the reduced order model is updated. The loop is<br />

executed until a residual drop of four to five orders of magnitude is obtained.<br />

This new technique for FSI problems has proved to be robust and powerful, and<br />

is ideally suited for FSI problems with strong coupling. <br />

References:<br />

1 Vierendeels, J.A., Riemslagh, K., Dick, E. and Verdonck, P.R.: Computer Simulation of<br />

Intraventricular Flow and Pressure Gradients During Diastole. Journal of Biomechanical<br />

Engineering – Transactions of the ASME, 122(6):667-674, 2000.<br />

2 Vierendeels, J.A.: Implicit Coupling of Partitioned Fluid-structure Interaction Solvers using<br />

a Reduced Order Model. In Proc. of the 35th AIAA Fluid Dynamics Conference and<br />

Exhibit, June 6-9 2005, Toronto, Ontario, Canada. AIAA-2005-5135. AIAA Meeting<br />

papers on disc, vol. 10, nr 11-12. ISBN 1-56347-763-7.<br />

10 <strong>Fluent</strong> News · Spring 2006


FLUID-STRUCTURE INTERACTION<br />

FSI Controls<br />

Flow Rate<br />

By Subham Sett, ABAQUS, Inc., Providence, Rhode Island, USA<br />

FLOW CONTROL DEVICES are commonly employed in<br />

the automotive, biomedical, and consumer appliances<br />

industries to maintain a constant bulk flow rate for varying<br />

inlet pressures. Fluid pressure variations may result<br />

from pipe friction loss, downstream restrictions, distance<br />

from the water tower, or elevation of the water tap, for<br />

example. Minimizing the impact of inlet pressure variation<br />

on the flow is essential for the proper performance of<br />

fluid systems.<br />

A schematic cross<br />

section of a typical<br />

VernaFlo device<br />

before (top) and<br />

after (bottom) the<br />

rubber component<br />

is deformed<br />

Vernay VernaFlo ® flow controls are custom-designed flow<br />

management devices used in a wide range of applications<br />

where consistent, reliable fluid flow is essential. This study<br />

evaluates the performance of a custom VernaFlo device,<br />

using a fully coupled fluid-structure interaction (FSI)<br />

analysis with ABAQUS and FLUENT. The computational<br />

results compare favorably with available experimental<br />

data.<br />

In a typical VernaFlo device, an elastomeric rubber component<br />

is housed inside the flow path. This rubber insert<br />

rests on a rigid seat and deforms under the influence of<br />

the incoming flow. At low operating pressures the rubber<br />

component undergoes very little deformation and allows<br />

flow<br />

The CFD sub-domain where the fluid flows (red) and the ABAQUS sub-domain<br />

which, for the VernaFlo valve, is a rubber insert that deforms for different flow rates<br />

(white)<br />

Static flow pressure (top) and Mises stress (bottom) at 20, 40, 60, and 90 psi; at high pressures, the rubber insert moves closer to the axis,<br />

stabilizing the flow rate<br />

<strong>Fluent</strong> News · Spring 2006 11


FLUID-STRUCTURE INTERACTION<br />

2.5<br />

Flow rate (L/min)<br />

2.0<br />

1.5<br />

1.0<br />

0.5 ABAQUS-FLUENT<br />

Experiment<br />

0.0<br />

0 20 40 60 80 100<br />

Inlet pressure (psi)<br />

120<br />

Experimental validation of the computation results<br />

the flow to develop. With increasing upstream pressure, the deformation increases, restricting<br />

the orifice diameter and thus limiting the fluid flow. Capturing the interaction between<br />

the fluid flow and the structural deformation is critical to accurately predicting the shape of<br />

the rubber component and the subsequent flow behavior.<br />

The FSI analysis of the valve with ABAQUS and FLUENT makes use of a co-simulation technique<br />

managed by MpCCI (Mesh-based parallel Code Coupling Interface). The fluid and<br />

structural domains are modeled and solved separately, and solution information is<br />

exchanged at the fluid-structure interface. The ABAQUS sub-domain includes the rubber<br />

component, which is modeled using reduced order hybrid axisymmetric elements. The<br />

rubber component presses against a rigid seat. Penalty contact with a friction coefficient of<br />

0.5 is defined between the rubber and the rigid seat. The simulation includes the effect<br />

of geometrical and material nonlinearities. The CFD sub-domain models the flow path,<br />

including a short upstream section, the section around the rubber component, and a long<br />

downstream section. The upstream variation in pressure is accounted for using a pressureinlet<br />

boundary condition, and a zero gauge pressure is used at the outlet. To enable local<br />

remeshing, the flow path is modeled using triangular elements. The water is modeled as a<br />

turbulent, incompressible fluid, and the cavitation model is enabled.<br />

Velocity magnitude contours on a 2D model of the<br />

cross flow fan; the flow is from bottom to top<br />

The fluid-structure interface is defined using MpCCI. During the simulation, the pressures<br />

acting on the rubber component in the fluid sub-domain are mapped and transferred to the<br />

structural sub-domain in ABAQUS via MpCCI. ABAQUS computes the deformations and the<br />

resulting stress state in the structure. The interface deformation quantities are then mapped<br />

and transferred from the structural sub-domain to the fluid sub-domain in FLUENT via MpCCI.<br />

This process of exchanging solution quantities continues incrementally until the analysis is<br />

complete.<br />

The analysis results identify how variations in inlet pressure affect the bulk fluid flow rate<br />

through the device, and how the fluid flow affects the deformation of the rubber component.<br />

Contour plots of the fluid pressures at inlet pressure levels of 20, 40, 60 and 90 psi and<br />

the corresponding deformed shapes show that with an increase in the inlet pressure, the rubber<br />

component deforms and undergoes increased contact with the rigid seat. Partial contact<br />

is observed at 20 psi, and full contact is established at 60 psi. The constriction path narrows<br />

during the inlet pressure rise, resulting in an increased resistance to the fluid flow and a<br />

region of very high stresses. At a given deformation state, the increased material stiffness<br />

helps maintain the bulk flow rate at a fairly constant level. As the flow quickens through the<br />

narrow constriction region, the fluid pressure drops significantly, resulting in a dramatic drop<br />

in the absolute pressure of the liquid. Cavitation occurs at higher upstream pressures.<br />

Contours of stream function for the cross flow fan<br />

The computational flow rate results were found to be in good agreement with experimental<br />

data. The flow rate increases from 0 to 2.1 liters/min during the initial pressure ramp-up from<br />

0 to 20 psi and has nearly constant flow in the operating pressure range of 20 to 120 psi. <br />

Acknowledgments<br />

ABAQUS, Inc. would like to thank Jim Bailey at Vernay Labs for providing the model and experimental<br />

data for the Vernay VernaFlo device, and David Schowalter at <strong>Fluent</strong> Inc. for assisting with the cavitation<br />

modeling.<br />

Velocity vector detail shows recirculation regions in<br />

one area of the cross flow fan<br />

12 <strong>Fluent</strong> News · Spring 2006


HVAC<br />

Fan Research Makes<br />

Cool Air Conditioners<br />

By Teng-Kiat Lim Ph.D. and Chee-Onn Chan, Department of Research and Application, O.Y.L. R&D Center, Selangor, Malaysia<br />

SIGNIFICANT TECHNICAL CHALLENGES face<br />

companies in the HVAC industry today. In an<br />

increasingly competitive global market, HVAC systems<br />

need to make use of advanced technologies<br />

and improved component performance. At the<br />

O.Y.L. R&D Center, CFD is used for the analysis and<br />

optimization of fans, compressors, heat exchangers,<br />

and ducts, and it plays an important role in HVAC<br />

product lifecycle management (PLM) as well. CFD<br />

helps reveal flow structures and comprehensive flow<br />

field information where experimental techniques<br />

cannot provide adequate resolution.<br />

Consider, for example, a split-type air conditioner<br />

that consists of one outdoor condensing unit (housing<br />

an air-cooled condenser, propeller fan, and<br />

refrigerant compressor) and one or more indoor<br />

units (housing an evaporating coil and cross flow<br />

fan). Split-type air conditioners are popular in residential<br />

buildings because of their simplicity and<br />

flexibility. If the air flow in the outer unit is not<br />

properly controlled, however, poor recirculation can<br />

cause the coil temperature to rise and the system<br />

performance to suffer. A series of simulations has<br />

been performed to study the various components<br />

of this type of system, and based on the results,<br />

modifications have been made.<br />

A 3D model of a 3-bladed fan in the outdoor condensing<br />

unit was created and used to visualize the<br />

air flow near the housing and in the discharge. The<br />

steady multiple reference frames (MRF) model in<br />

FLUENT was used along with the standard k-ε<br />

turbulence model and standard wall functions. A<br />

tetrahedral mesh of about 1 million cells was generated<br />

using GAMBIT and TGrid. The triangular surface<br />

mesh was refined in the regions around the<br />

hub, tip, and blade edges. A uniform air velocity<br />

was imposed at the condenser inlet boundary, and<br />

a constant pressure was applied to the outlet. The<br />

results illustrated that when the steady incoming air<br />

flow enters the rotating zone of the fan, a strong<br />

radial component develops, driving the flow<br />

towards the side walls. This behavior is enhanced by<br />

the existence of an outlet front panel that partially<br />

blocks the flow. The CFD predictions were found to<br />

be in good agreement with air flow test measurements.<br />

Studies on new front panel designs are<br />

currently being conducted so that better air flow<br />

performance can be achieved.<br />

A two-bladed fan has also been studied. For this<br />

case, an unstructured mesh of about 1.2 million elements<br />

was needed to give a mesh-independent<br />

result. The 3D, double precision simulation also<br />

made use of the MRF and standard k-ε models. The<br />

results indicated that even though the front panel<br />

impacts the fan performance, a more uniform air<br />

flow distribution can be achieved, along with a<br />

lower pressure drop.<br />

The cross flow fan (CFF) operates in a fundamentally<br />

different way than axial or centrifugal fans. Flow<br />

enters across the full width of the rotating impeller<br />

and exits on the opposite side. The flow structure<br />

inside the impeller can be divided into two regions:<br />

an eccentric vortex or recirculation region having<br />

closed streamlines, and a through-flow region connecting<br />

the inflow and outflow sectors. The blades<br />

typically have a uniform thickness and a circular arc<br />

profile. The performance and stability of a CFF are<br />

governed by the geometrical parameters of the<br />

impeller and casing. Using the sliding mesh model,<br />

a transient 2D analysis was performed to study the<br />

effect of parameters such as blade angle, radius<br />

ratio, and number of blades. Maximum energy<br />

transfer through the impeller takes place in the<br />

region where the flow follows the blade curvature.<br />

Radial velocity is not uniform through the blade<br />

channels. More flow leakage is observed through<br />

the tongue clearance portion at low flow coefficients<br />

and static pressure is always negative in and<br />

around the impeller region <br />

Absolute pressure on the 3-bladed fan (top) and<br />

2-bladed fan (bottom)<br />

<strong>Fluent</strong> News · Spring 2006 13


AUTOMOTIVE<br />

Citroën<br />

Vortex structures colored by streamwise velocity show<br />

the instantaneous and highly three-dimensional flow;<br />

note the interaction between the streamwise vortices<br />

formed at separation on the front fender and the<br />

A-pillar on the leeward side<br />

Postprocessed by Ensight from CEI and Distene<br />

AS PASSENGER VEHICLES BECOME LARGER, their stability<br />

and safety characteristics become more sensitive to aerodynamic<br />

forces. Such effects are particularly important for vehicles exposed<br />

to a crosswind [1]. For instance, all drivers are familiar with the<br />

effect of cross-wind gusts, like those that may occur when a vehicle<br />

emerges from a tunnel or during overtaking. CFD can provide<br />

detailed information for understanding the changes in car behavior,<br />

handling, and performance under these conditions.<br />

Simulations are routinely performed at PSA using engineering<br />

approaches that are based on the solution of the Reynolds-averaged<br />

Navier-Stokes (RANS) equations. While this approach is often<br />

adequate for steady attached flows with no recirculation regions,<br />

numerous studies have shown the limitations of the RANS<br />

approach to accurately predict flows with massive separations and<br />

that are fundamentally unsteady.<br />

Large eddy simulation (LES) can have greater success than RANS<br />

methods in predicting separation and large scale unsteady flows,<br />

particularly in the wake and on the leeward side of vehicles under<br />

crosswind conditions. In recent years, this type of unsteady<br />

calculation has been increasingly used in industry and FLUENT has<br />

made a special effort to develop industrial-strength validation<br />

examples for LES modeling techniques [2,3].<br />

PSA Peugeot Citroën has developed a strong expertise in CFD for<br />

external aerodynamics, cabin modeling, underhood thermal management,<br />

and more. In-house methodologies have been developed<br />

and are regularly improved, in order to fulfill new project<br />

requirements, which are getting more and more demanding, both<br />

in terms of accuracy and turn-over time. Computations are done<br />

in conjunction with wind tunnel experiments, either on full<br />

scale prototypes or, as in the case presented here, on a 1:5 scale<br />

model. Different measurement techniques are used, such as<br />

particle-image velocimitry (PIV), pressure taps, and tomography.<br />

The different methodologies also benefit from a collaboration<br />

between PSA and <strong>Fluent</strong>, and this was the basis of a recent joint<br />

project to address the potential benefit and cost of LES compared<br />

to the current RANS approach for semi-realistic configurations.<br />

Fujitsu/Siemens provided access to a large-scale high performance<br />

computing (HPC) facility for this project.<br />

Model Exp SST k-ω LES WALE LES WALE RSM v 2 -f<br />

(35M cells) (65M cells)<br />

Drag (SCx) 0.70 0.66 0.69 0.68 0.71 0.73<br />

14 <strong>Fluent</strong> News · Spring 2006<br />

Detailed force and torque components<br />

acting on the vehicle; under crosswind<br />

conditions, the side force (SCy) and<br />

yaw moment (SCn) are the most<br />

influential for vehicle stability<br />

Forces<br />

Moments<br />

Side (SCy) 2.22 2.00 2.19 2.18 2.30 2.10<br />

Lift (SCz) 1.40 1.66 1.27 1.30 1.82 1.77<br />

Yawing (SCn) -0.64 -0.60 -0.57 -0.59 -0.47 -0.47<br />

Rolling (SCl) -0.42 -0.36 -0.49 -0.49 -0.46 -0.41<br />

Pitching (SCm) 0.12 0.10 0.21 0.23 0.03 0.07


AUTOMOTIVE<br />

C5 in a Crosswind<br />

By Sylvain Lardeau, PSA Peugeot Citroën Automobile, Vélizy-Villacoublay, FRANCE, and Fabrice Mathey and Nicolas Vallée, <strong>Fluent</strong> France<br />

The RANS and LES simulations were performed on a<br />

1:5 scale simplified model of a Citroën C5. The crosswind<br />

effect was produced with the model placed at a<br />

yaw angle of 20°. The Reynolds number of the flow,<br />

based on the incoming velocity and car height, H,<br />

was 6.19 x 10 5 . The surface mesh (around 730,000<br />

elements) was generated with ANSA and was locally<br />

refined close to the edges and on the leeward faces of<br />

the vehicle. The boundary layers were generated with<br />

the preprocessor TGrid, using 5 layers of prisms. The<br />

remaining part of the domain consisted of tetrahedral<br />

cells and was generated with the preprocessor<br />

GAMBIT. Sizing functions were used to control the<br />

growth rate and the cell size of the mesh in the region<br />

of separation on the leeward side of the car. The<br />

minimum resolution required to resolve the relevant<br />

turbulent length scales was estimated from prior RANS<br />

simulations. Two grids (of 35 million and 65 million<br />

cells) were created to address the sensitivity of the<br />

results to the mesh resolution. The body was placed in<br />

an open channel with a cross section of 14.5H x 7H.<br />

The channel inlet was located 5H from the front face<br />

of the body, and the channel outlet was located 8.5H<br />

from the rear. The boundary layers were modeled with<br />

the Werner-Wengle wall-function approach. The average<br />

y + for the first cell off the wall was less than 50.<br />

The simulation benefited from the recent enhancements<br />

in numerics and sub-grid scale (SGS) modeling<br />

in FLUENT 6.2. The bounded central differencing<br />

scheme was used to prevent unphysical wiggles<br />

numerically introduced by the pure central differencing<br />

scheme. The non-iterative time advancement, or<br />

NITA algorithm significantly increased the speed of<br />

the transient calculation. The wall-adapting local<br />

eddy-viscosity (WALE) SGS model, which returns the<br />

correct wall asymptotic variation of the turbulent<br />

viscosity in a cost-effective manner, was used as<br />

well. The simulations were performed on a 96 CPU<br />

Fujistu cluster. For the 35M case, statistics were<br />

gathered during 16,000 iterations for an elapsed<br />

time of 70 hours. Statistics were gathered during<br />

6,000 iterations for the 65M case, for an elapsed<br />

time of 50 hours.<br />

The full mean aerodynamic forces and torques were<br />

compared with RANS methods and experimental<br />

data. Under crosswind conditions, the side force and<br />

yaw moment are the most influential for vehicle stability.<br />

The LES results showed very good agreement<br />

with the experimental data for these quantities. The<br />

difference between measurements and calculations<br />

was 2% for the side force coefficient (SCy), and was<br />

in the range of 10% for the yawing moment (SCn).<br />

The LES runs also gave consistent results for the<br />

other aerodynamic force components; the drag<br />

force (SCx) was predicted with a 1% error, for<br />

instance. It is noteworthy that even if some RANS<br />

simulations are able to predict side force and yawing<br />

moment with reasonable accuracy, none of the turbulence<br />

models could predict consistent results for<br />

all of the force components and moments. Surface<br />

pressure coefficients also confirm the differences<br />

between the different computational methods used.<br />

On the A-pillar on the leeward side, especially, the<br />

pressure drop is too strong with the RANS model.<br />

The two LES runs gave comparable results, indicating<br />

that reasonable mesh independence was<br />

reached with the 35 M grid, as far as global force<br />

coefficients are concerned.<br />

The visualization of the mean velocity streamlines<br />

indicates the formation of two streamwise vortices on<br />

the leeward side of the vehicle. The first one is formed<br />

on the front fender and the second is formed on the<br />

A-pillar and along the side window. Animation and<br />

instantaneous flow visualization demonstrate the<br />

highly three-dimensional, unsteady nature of these<br />

vortices. The predominant peak frequencies calculated<br />

from a Fourier transform of the pressure signal at<br />

several locations on the leeward side is equal to<br />

110 Hz, a value very close to the experimental peak<br />

frequency of 100 Hz.<br />

These simulations demonstrate how RANS and LES<br />

modeling approaches can complement each other<br />

to study car aerodynamics. Current RANS-based<br />

methodologies are able to correctly predict drag<br />

force components for attached flow and mild separation<br />

on realistic car geometries under uniform wind<br />

conditions. LES is a better tool for studying vehicle<br />

aerodynamics in crosswind conditions, since it provides<br />

a more accurate drag force component and<br />

pressure distribution along the vehicle. In addition,<br />

LES gave insight into the unsteady behavior of the<br />

flow, an important feature for both vehicle stability<br />

studies and aeroacoustic applications. <br />

References:<br />

1 Ryan, A. and Dominy, R.G.: The Aerodynamic Forces<br />

Induced on a Passenger Vehicle in Response to a<br />

Transient Cross-wind Gust at a Relative Incidence of<br />

30 Degrees. SAE Paper 980392, 1998.<br />

2 Kim, S.E.: Large Eddy Simulation using Unstructured<br />

Meshes and Dynamic Subgrid-scale Turbulence Models.<br />

AIAA Paper no. 2004-2548, 2004.<br />

3 Mathey, F. and Cokljat, D.: Zonal Multi-domain<br />

RANS/LES Simulation of Air Flow over the Ahmed Body.<br />

In: Engineering Turbulence Modelling and Experiments,<br />

6, pp. 647-656, Elsevier Ltd., 2005.<br />

0<br />

-0.5<br />

-1.0<br />

Cp<br />

-1.5<br />

-2.0<br />

Experiment<br />

LES<br />

RSM high-Re<br />

v 2 -f<br />

SST k-ω<br />

-2.5<br />

-1.0 -0.5 0.0 0.5 1.0 1.5 2.0<br />

X (m)<br />

Surface mesh of the vehicle generated with ANSA<br />

Surface pressure distribution on the windward side on<br />

the “eye plane”, z=1.250 m<br />

<strong>Fluent</strong> News · Spring 2006 15


AUTOMOTIVE<br />

TORQUE CONVERTERS ARE FLUID-COUPLING DEVICES that<br />

transmit power from vehicle engines to the wheels. They are used<br />

in automobiles with automatic transmissions to smoothly control<br />

the torque supplied to the wheels at all speeds.<br />

An exploded view of the torque converter showing<br />

the pump (top), stator (middle), and turbine (bottom)<br />

Torque<br />

Converters<br />

Get In Gear<br />

By Vamshi Korivi, DaimlerChrysler, Auburn Hills, Michigan, USA<br />

Sandeep Sovani and Pepi Maksimovic, <strong>Fluent</strong> Inc.<br />

Torque converters are essentially chambers filled with transmission<br />

fluid, with the primary active parts being the pump (or impeller),<br />

stator, and turbine. The impeller blades rotate at the engine<br />

speed and impart angular momentum to the fluid as they force it<br />

radially outward. The turbine blades receive the fluid from the<br />

pump and turn it radially inward. The angular momentum thus<br />

absorbed by the turbine is passed to the transmission as torque.<br />

The flow exits from the turbine blade passages with little remaining<br />

angular momentum and encounters the stator blades, which<br />

accelerate the fluid back to the pump rotational speed. This action<br />

creates a torque on the stator, which is arrested by a non-rotating<br />

shaft built into the transmission housing. Torque multiplication<br />

results from the fact that the turbine torque is larger than that of<br />

the pump.<br />

Numerical simulations of torque converters are important because<br />

the complex topology and rotation of these devices make detailed<br />

experimental investigation virtually impossible. Simulations help<br />

identify optimal designs that yield improved performance and fuel<br />

economy. Using the multiple reference frames (MRF) model in<br />

FLUENT, a DaimlerChrysler torque converter has been studied.<br />

The MRF model allows the different parts of the torque converter<br />

(the pump, stator, and turbine) to be simulated using one relative<br />

orientation, but separate rotating frames. The model is well suited<br />

for torque converters, since the number of blades is chosen to<br />

prevent the amplification of harmonics and possible structural<br />

damage to the parts. This means that any single relative position<br />

of the blades is a good representation of the averaged behavior of<br />

the system as a whole. For the torque converter considered here,<br />

the number of blades for the pump, turbine, and the stator is 31,<br />

29, and 20 respectively.<br />

Contours of total pressure on the blade surfaces of<br />

the three elements; large pressure losses can be seen<br />

across the stator blades (in the middle)<br />

K-factor<br />

A hexahedral mesh of about 2.9 million cells was used for the initial<br />

calculation. A pressure gradient-based mesh adaption was<br />

subsequently performed to enhance the solution accuracy. Using<br />

transmission fluid and a pump speed of 2000 rpm, the turbulent<br />

flow was captured using the realizable k-ε turbulence model.<br />

Custom postprocessing tools were created for the automatic<br />

extraction of torque converter quantities of interest, such as the<br />

mass-averaged flow angles at the inlet and outlet of each element,<br />

the efficiency, and the torque imbalance, for example.<br />

The results illustrated several flow features. Total pressure contours<br />

were used to illustrate the pressure loss across the stator blades.<br />

Velocity vectors were used to check for separation regions. The<br />

k-factor, the ratio of pump torque to the square root of pump<br />

speed and an indicator of the torque converter capacity, was<br />

found to be within 2% of measured values at all speed ratios (the<br />

ratio of turbine speed to pump speed) studied.<br />

Efficiency<br />

FLUENT<br />

Experiment<br />

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8<br />

Speed ratio<br />

Predictions of efficiency and k-factor as a function of<br />

speed ratio show good agreement with measured values<br />

0.9<br />

Predictions of efficiency, defined as the product of torque ratio<br />

and speed ratio, were found to be within 1.6% of the data at a<br />

speed ratio of 0.7. The torque ratio, defined as the ratio of turbine<br />

torque to pump torque, was found to be within 5% for speed<br />

ratios above 0.2. The ability to predict efficiency and torque ratios<br />

at high speed ratios is important for improving fuel economy. <br />

16 <strong>Fluent</strong> News · Spring 2006


AUTOMOTIVE<br />

Emissions Control Through<br />

Carbon Canisters<br />

By Ranjit Singh, Expert Corporation, and Pepi Maksimovic, <strong>Fluent</strong> Inc.<br />

CARBON CANISTERS ARE DEVICES commonly<br />

used to control the emissions of volatile hydrocarbons.<br />

Hydrocarbons (HC) are hazardous to human<br />

health and the environment. For automobiles, HC<br />

emissions are produced during the filling of the fuel<br />

tank, and during vehicle operation. When the<br />

engine is off, evaporation from the vehicle fuel<br />

system still occurs, even at ambient temperatures<br />

typical of the diurnal cycle. Allowable HC emission<br />

limits are set by government regulations; for example,<br />

the LEV II (Low Emitting Vehicle-II) standard<br />

allows a certain amount of hydrocarbon emissions<br />

for a specific range of gross vehicle weight.<br />

Carbon canisters are part of the evaporative emission<br />

control system, which includes the fuel tank,<br />

vent and purge valves, and fuel lines. The role of the<br />

carbon canister is to store the fuel vapor generated<br />

in the system instead of having it escape into the<br />

atmosphere. The HCs are then burned off by purging<br />

the canister into the intake manifold when the<br />

engine is running. An optimum design includes a<br />

high working capacity, minimal pressure drop in the<br />

evaporative emissions system, space and size restrictions,<br />

and the ability to meet the mandated emission<br />

standards. Furthermore, the carbon utilization<br />

in the canister should be uniform; when less carbon<br />

material is needed, savings are realized. In today’s<br />

competitive market, carbon canister manufacturers<br />

must be able to predict canister performance without<br />

having to build and test various prototypes.<br />

desorbing), until breakthrough is achieved, i.e.<br />

when a set cumulative amount of butane passes<br />

through the canister outlet. Arrhenius reaction rates<br />

for the adsorption and desorption reactions were<br />

prescribed. The carbon pellets were modeled as<br />

a porous region. A mesh of 184,000 hexahedral<br />

elements was used for the laminar simulation.<br />

The CFD results were used to assess the design.<br />

Using pathlines, a region with little or no flow was<br />

identified in the first chamber. Ideally, the canister<br />

design should provide uniform utilization of the<br />

carbon media in all of the chambers, and this result<br />

suggests that it does not. Mass fraction contours<br />

were used to illustrate the distribution of butane at<br />

several times during the canister loading process.<br />

The growth of butane inside the canister is the<br />

result of the ongoing surface reactions with the carbon<br />

pellets, so this provides another measure of<br />

how well the unit is operating as a whole. A plot of<br />

the butane mass fraction at the outlet as a function<br />

of loading time can also be used to indicate the<br />

time when breakthrough is achieved. The carbon<br />

canister capacity is computed by integrating under<br />

this curve. <br />

The geometry of the carbon canister, showing the<br />

inflow port (blue), 3 chambers packed with carbon<br />

pellets (gray) and vent port (yellow)<br />

Pathlines show a dead zone in the upper left corner<br />

in the first carbon-filled chamber; the presence of<br />

dead zones points to underutilization of carbon<br />

One carbon canister design from Expert Corp. has<br />

been simulated using FLUENT. The purpose of the<br />

CFD simulation was twofold: first, to compute the<br />

canister capacity and pressure drop at 60 l/min of<br />

air flow, and second, to predict canister breakthrough<br />

for a gaseous mixture of butane and nitrogen<br />

entering the canister at 15 g/hr. The mixture<br />

reacts with packed carbon pellets (adsorbing and<br />

Contours of butane mass fraction on a slice through the canister during loading (left) and once the canister has<br />

reached capacity (right)<br />

<strong>Fluent</strong> News · Spring 2006 17


POWER TOOLS<br />

Scavenging in a Stratified, Charged<br />

Two-Stroke Engine<br />

By Wolfgang Emmerich, SOLO Kleinmotoren GmbH, Sindelfingen, Germany<br />

TWO-STROKE GAS ENGINES are commonly used in hand-held machines,<br />

such as chain saws and grass trimmers. These engines have a tendency to cause<br />

pollution as a result of unburned hydrocarbons being released along with the<br />

exhaust. At SOLO Kleinmotoren, a new by-pass system has been developed that<br />

allows the combustion chamber to be flushed with either a lean fuel mixture or<br />

pure air early in the engine cycle. Fuel is injected as a rich mixture later in the<br />

cycle when the outlet port is closed. The by-pass system prevents the fuel from<br />

escaping from the combustion chamber before combustion is complete.<br />

A numerical investigation of the by-pass system has been carried out using the<br />

dynamic mesh model in FLUENT. Throughout several engine cycles, the mass<br />

fractions of pure air, a rich fuel mixture, and a mixture of exhaust gases were<br />

tracked. Combustion was approximated by resetting the pressure, temperature,<br />

and components of the three gaseous species each time the piston was in the top<br />

dead center (TDC) position.<br />

The operation of the engine is illustrated at right. During the suction phase, clean<br />

air is drawn into the crank case (blue arrow) and a fresh mixture of fuel and air<br />

is drawn into the by-pass channel (green arrow). As the crank case compresses,<br />

the clean air enters the combustion chamber through transfer ports, while the<br />

fuel mixture is pressed into the chamber through an injection port. Due to fluid<br />

dynamic constraints dictated by the local flow, the fresh air in the chamber prevents<br />

the fuel mixture from leaving the combustion chamber directly through<br />

the outlet port. Only exhaust gases (red) and pure air can be released into the<br />

environment.<br />

A schematic of the engine with the by-pass system: pure air (blue) and fuel (green)<br />

enter through ports on the left; following combustion, pure air and exhaust gases<br />

leave through the outlet port on the right<br />

The distributions for the three gaseous species at three different piston positions<br />

illustrate the success of the by-pass system. The blue surface corresponds to fresh<br />

air, the green surface to the fuel mixture, and the red surface to the exhaust gas<br />

mixture. At bottom dead center (BDC), the positions of the surfaces indicate that<br />

the fresh air forms a fluid shield between the fuel mixture and the outlet port.<br />

When the piston has advanced 50° beyond BDC, the fluid shield is not perfect,<br />

so a small portion of the fuel mixture can leak past the shield near the cylinder<br />

walls. When the piston has advanced 100° after BDC, the outlet port is closed by<br />

the rising piston, just before the fuel mixture has a chance to escape.<br />

The results of the simulation show the ability of the new engine to reduce environmental<br />

pollution by limiting the release of unburned hydrocarbons. The<br />

results also suggest that there is potential to improve the fluid dynamics of the<br />

fuel injection as well as the flushing of the combustion chamber to further<br />

enhance the performance of the new by-pass system. <br />

Surfaces of fresh air (blue), fuel (green), and exhaust (red) at three piston positions:<br />

a) bottom dead center (BDC), where the air forms a fluid shield to block the fuel,<br />

b) 50° beyond BDC, where some fuel leaks past the air shield near the cylinder wall,<br />

and c) 100° beyond BDC, at which time the rising piston has closed the outlet port<br />

18 <strong>Fluent</strong> News · Spring 2006


F L U E N T N E W S S U P P L E M E N T<br />

Focus on CFD<br />

For the Process Industries<br />

s2 OVERVIEW<br />

Transporting CFD to<br />

Process Engineers<br />

s3 MIXING<br />

Tracing Homogenization<br />

s5<br />

Finding the Optimum<br />

Blend Time Calculation<br />

s7 MICROREACTORS<br />

Liquid Mixing<br />

in Microreactors<br />

s8<br />

THERMAL RUNAWAY<br />

Preventing Runaway<br />

Reaction Accidents<br />

s10 EMISSIONS<br />

Scrubbers for<br />

Flue Gas Cleanup<br />

s12 FURNACES<br />

Ultra-low NOx Burners<br />

Get Cracking<br />

s13 MULTIPHASE<br />

Controlling Droplet Size<br />

Distribution in Emulsions<br />

s14 Understanding<br />

Fluid-Bed Coating<br />

s15 PUMPS<br />

Pumping out New<br />

Designs More Quickly<br />

s16 EXTRUSION<br />

Polymer Processing<br />

Simulation for Foam<br />

Extrusion


OVERVIEW<br />

PROCESS INDUSTRIES<br />

Transporting CFD<br />

to Process Engineers<br />

By Ahmad Haidari, Chemicals and Process Industry Segment Manager, <strong>Fluent</strong> Inc.<br />

CHEMICAL AND OTHER PROCESS INDUSTRY<br />

companies are driving issues such as green engineering,<br />

energy reduction, chemical sustainability and corporate<br />

profitability through innovation and the use of ingenious<br />

engineering and technology. Simulation software plays a<br />

key role in helping engineers to better understand<br />

processes and is assisting in productivity and efficiency<br />

gains across the industry. CFD in particular is gaining<br />

broad traction with researchers, process engineers, and<br />

equipment designers to help analyze and design the flow<br />

and performance of process equipment such as stirred<br />

tanks, fluidized bed reactors, separators, combustion<br />

systems, heat exchangers, and polymer and material<br />

processing and handling equipment.<br />

formulation accounts for phenomena such as nucleation,<br />

growth, dispersion, dissolution, aggregation, and<br />

breakage, so it is possible to describe and track changes<br />

in the particle population. Applications of this technology<br />

span a wide range of multiphase systems such as<br />

solid-liquid dispersions, crystallization, precipitative reactions,<br />

gas-sparged reactors, liquid-liquid dispersions, and<br />

liquid-liquid separation equipment. Second, for granular<br />

mixtures, the macroscopic particle model (MPM), available<br />

through user-defined functions (UDFs) in FLUENT,<br />

allows for the presence of large particles where particle<br />

collisions, rotation, adhesion, and other forces are<br />

accounted for using a hard sphere approach. Third,<br />

modifications to the volume of fluid (VOF) free surface<br />

model in FLUENT have been developed to simulate capillary-driven<br />

flow in an unsaturated porous region.<br />

The primary modeling goal was to study the spread<br />

of unbound liquid inside the voids formed by fibrous<br />

material, but the model also accounts for sorption (the<br />

physical bonding of moisture to the fiber), since the<br />

performance of many hygiene, food, and consumer<br />

products strongly depends on their sorption capacity.<br />

The flow of macroscopic particles<br />

in a hopper<br />

Technical advances at <strong>Fluent</strong> have progressed in two<br />

distinct directions. The traditional focus on more<br />

reliable and advanced physical models is being complemented<br />

by highly customized interfaces to fit the<br />

modeling needs of a specific application or equipment<br />

design. These advances, combined with the successful<br />

track record of CFD in many companies, have brought<br />

the full power of CFD to process engineers.<br />

The applicability of CFD to process industry applications<br />

is achieved through ongoing focused development<br />

and user-driven projects. For example, three customerdriven<br />

technology developments have been implemented<br />

recently that uniquely expand FLUENT’s capabilities.<br />

First, a population balance modeling effort was driven<br />

by a need to combine detailed fluid mechanics analysis<br />

with a better description of the size distribution of particles,<br />

bubbles, or droplets in a multiphase mixture. The<br />

In this special newsletter supplement, a diverse sampling<br />

of process industry applications is presented.<br />

Traditional applications such as mixing in stirred tanks<br />

are joined by mixing in microchannels. Reacting flow is<br />

covered in the context of thermal runaway (and a<br />

promising antidote), the cleanup of emissions, and<br />

the optimization of an ethylene cracking furnace.<br />

Multiphase flow examples include a fluidized bed and<br />

droplet breakup validations. The use of CFD for designing<br />

special purpose pumps is also reviewed. Finally, simulations<br />

of the extrusion of foam – a gas-impregnated<br />

polymer – are described.<br />

As these articles illustrate, the use of CFD is spreading<br />

to more and more applications within the process<br />

industries. With increases in computing capacity and<br />

the advances in technology currently taking place, this<br />

expansion can only continue in the years to come. <br />

<strong>Fluent</strong> News is published by<br />

10 Cavendish Court • Lebanon, NH 03766 USA<br />

603 643 2600 • www.fluent.com<br />

© 2006 <strong>Fluent</strong> Inc. All rights reserved.<br />

Editor: Liz Marshall<br />

Assistant Editor: Susan Wheeler<br />

Contributing Editors: Erik Ferguson and Keith Hanna<br />

Design: Lufkin Graphic Designs<br />

An iso-surface of liquid at two times, corresponding to a 10% saturation level<br />

predicted by FLUENT’s wicking model<br />

FLUENT, FiDAP, GAMBIT, POLYFLOW, G/Turbo, MixSim,<br />

FlowLab, Icepak, Airpak, and FloWizard are trademarks of<br />

<strong>Fluent</strong> Inc. Icepak and Airpak are joint developments of<br />

<strong>Fluent</strong> Inc. and ICEM-CFD Engineering. All other products<br />

or name brands are trademarks of their respective holders.<br />

S2 <strong>Fluent</strong> News · Spring 2006


MIXING<br />

The two-impeller system with conductivity probes<br />

Contours and vectors of velocity magnitude in a<br />

vertical plane mid-way between two baffles,<br />

computed by the sliding mesh model and the<br />

standard k-ε (top) and LES (bottom) turbulence<br />

models<br />

Tracing<br />

Homogenization<br />

By Michal Mo˘stĕk and Milan Jahoda, Department of Chemical Engineering,<br />

Prague Institute of Chemical Technology, Prague, Czech Republic<br />

THE BLENDING OF MISCIBLE LIQUIDS is a<br />

common mixing operation in the chemical, biochemical,<br />

food-processing, and other industries.<br />

Most frequently, it takes place in mechanically<br />

agitated tanks, where the miscible liquids are<br />

blended to a predetermined degree of homogeneity.<br />

One of the critical parameters of a mixer<br />

is the time of homogenization, or blending time.<br />

In practice, it is widespread to use empirical correlations<br />

derived from experiments to predict<br />

mixing times. However, these correlations have<br />

many limitations, mainly in the design of<br />

non-standard tank geometries and in cases with<br />

multiple impellers.<br />

Using CFD, the study of processes taking place in<br />

mixing tanks becomes easier and more economical<br />

than with the use of experiments. Thanks to<br />

the growing performance of today’s computers<br />

and with the help of other timesaving tools like<br />

the preprocessor MixSim, an elementary simulation<br />

of a mixing system can be performed within<br />

an hour. For more advanced simulations, the possibility<br />

of parallel computations and the availability<br />

of new and faster non-iterative solvers today<br />

enable a detailed study of mixing tanks to be performed<br />

in a range of days and weeks, where it<br />

would have taken months not long ago.<br />

In mixing tanks, where baffles are used to eliminate<br />

the central vortex and prevent air entrainment<br />

in the system, two methods of simulating<br />

the impeller rotation are most often used. The<br />

multiple reference frames (MRF) method is a fast<br />

steady-state method, which together with the<br />

k-ε family of turbulence models, provides timeaveraged<br />

information about processes taking<br />

place in the system. The transient sliding mesh<br />

(SM) method is much more accurate and suitable<br />

for the simulation of homogenization. For both<br />

methods, the tank is divided into a cylindrical<br />

part containing the impeller and a stationary part<br />

containing the baffles and tank walls. In the MRF<br />

method the impeller maintains a fixed position<br />

relative to the baffles, and the solution in the<br />

impeller region is solved in the rotating frame of<br />

the impeller. For the SM method, the grid for the<br />

impeller region rotates in time, so that transient<br />

effects, especially those applied to phenomena<br />

such as blending, can be captured. While<br />

the time-averaged k-ε models are often used in<br />

combination with the SM model, a transient<br />

turbulence model can also be used for the most<br />

comprehensive simulation of a transient process.<br />

When this level of detail is of interest, a transient<br />

turbulence model such as large eddy simulation<br />

(LES) can be used. To get the most out of this<br />

kind of modeling effort in terms of accuracy and<br />

precision, care must be taken when setting up<br />

the model, in terms of the size and density of the<br />

computational grid, the time step, and a suitable<br />

choice of sub-grid scale turbulence model.<br />

Two examples of liquid homogenization in a<br />

tank filled with water and equipped with either<br />

one Rushton turbine or two 6-bladed 45°<br />

pitched blade turbines mounted on a common<br />

shaft are presented below. For both cases, the<br />

tank diameter, T, is 0.29m and the shaft rotates<br />

at a constant speed of 300rpm. A second liquid<br />

(a saturated solution of NaCl) is introduced to<br />

the tank and the progress of blending is studied.<br />

Conductivity probes are used for experimental<br />

observations of tracer concentration,<br />

and CFD models are run for comparison. The<br />

measuring volume of the probe is approximately<br />

0.55cm 3 . For the one-impeller system,<br />

the probe is placed between the baffles, with<br />

an off-bottom clearance of T/4 and at a distance<br />

of T/20 from the tank wall. For the twoimpeller<br />

system, one probe is placed near the<br />

tank bottom, with the same off-bottom clearance<br />

and distance from the wall as in the previous<br />

case. A second probe is placed at a distance<br />

of T above the first probe. The tracer is<br />

injected just below the free liquid surface, at a<br />

horizontal distance of T/4 from the tank wall,<br />

opposite the probe. Output signals from the<br />

PROCESS INDUSTRIES<br />

<strong>Fluent</strong> News · Spring 2006<br />

S3


MIXING<br />

PROCESS INDUSTRIES<br />

normalized concentration<br />

normalized concentration<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

0.0 2 4 6 8 10<br />

time (s)<br />

2.0<br />

1.5<br />

1.0<br />

experiments<br />

mrf ske<br />

sm ske<br />

sm les<br />

0.5<br />

experiments<br />

mrf ske<br />

sm ske<br />

sm les<br />

0.0<br />

0.0 5 10 15 20 25<br />

time (s)<br />

The time traces of normalized concentration<br />

recorded (top) at one location in the one-impeller<br />

tank and (bottom) at two locations in the twoimpeller<br />

tank during the homogenization process,<br />

comparing three modeling approaches with five<br />

identical experimental runs; the sliding mesh with<br />

LES approach is the closest fit to the data<br />

probes are processed by a conductivity meter,<br />

digitized by an A/D converter and registered<br />

by a computer for further processing in the<br />

form of dimensionless concentration values.<br />

In the simulations, structured grids composed of<br />

non-uniformly distributed hexahedral cells were<br />

made in GAMBIT. The tank volume was decomposed<br />

into many smaller volumes in order to<br />

make a grid usable for structured meshing. In<br />

the case of the one-impeller system, the whole<br />

vessel was divided into 611,000 hexahedral<br />

cells (99 x 37 x 168 along the axial, radial<br />

and tangential coordinates), in the case of the<br />

two-impeller system, the mesh consisted of<br />

1,200,000 hexahedral cells (198 x 37 x 168 along<br />

the axial, radial and tangential coordinates).<br />

The first group of simulations of the flow field was<br />

performed using the MRF technique. By tracking<br />

the mixing of two fluids with equal properties,<br />

the solution of a single species equation could be<br />

used to follow the mixing progress. Converged<br />

solutions were obtained and then the flow field<br />

was “frozen” and only the unsteady solution of<br />

the passive scalar equation was performed.<br />

In the case of the SM and standard k-ε (SKE)<br />

turbulence model, the MRF solution was used as<br />

a starting point. The simulation was then<br />

switched to unsteady, with a time step set to<br />

0.001s and the non-iterative time advancement<br />

(NITA) option. When the periodic steady-state<br />

was reached (after 25s in the case of the<br />

one-impeller system and about 40s in the case of<br />

the two-impeller system), the homogenization<br />

simulation was initiated in the same way as<br />

described for the MRF simulation. The only difference<br />

was that the computations of tracer distribution<br />

in the vessels were solved together with<br />

the flow equations as the impeller mesh continued<br />

to rotate.<br />

In the case of the SM and LES approach, the MRF<br />

solution was again used as a starting point. The<br />

dynamic Smagorinsky-Lilly sub-grid scale model<br />

was chosen to capture the effects of turbulent<br />

eddies smaller than a cell size. Special attention<br />

was also given to the discretization schemes used.<br />

In particular, the bounded central differencing<br />

(BCD) scheme was used for discretization of the<br />

momentum and species transport equations, the<br />

body force weighted scheme was used for pressure<br />

discretization, and the fractional step method was<br />

chosen for pressure-velocity coupling.<br />

Comparison of the velocity fields calculated using<br />

the standard k-ε and LES models show the<br />

increased detail provided by the LES approach.<br />

For the two-impeller system, concentration maps<br />

at different times of the LES computation illustrate<br />

the complex dispersion of the tracer during<br />

the blending process. For both vessels, comparisons<br />

of the calculated and experimental concentration<br />

profiles indicate that the sliding mesh<br />

approach using LES provides the closest overall<br />

match to the probe readings throughout the<br />

blending process. <br />

Contours and vectors of velocity magnitude in a<br />

vertical plane mid-way between two baffles,<br />

computed by the sliding mesh model and standard<br />

k-ε (left) and LES (right) turbulence models<br />

1 sec 5 sec 10 sec 15 sec<br />

Contours of normalized tracer concentration in a vertical plane mid-way<br />

between two baffles after the tracer addition<br />

S4 <strong>Fluent</strong> News · Spring 2006


MIXING<br />

Finding the Optimum<br />

Blend Time Calculation<br />

By Benoit Post, <strong>Fluent</strong> France, and Rajiv Lochan Rath, <strong>Fluent</strong> India<br />

THE MIXING OF SINGLE AND MULTIPHASE FLUIDS in stirred tank reactors<br />

is a common operation in many industries, such as chemicals, water treatment,<br />

pharmaceuticals, and petroleum. Understanding the fluid flow in these<br />

tanks is critical for equipment design, scale-up, process control, and economic<br />

factors. CFD is now being used routinely to provide this information,<br />

enabling engineers to select the best agitator design to obtain the desired<br />

process performance. Blending time evaluation is one of the key objectives of<br />

such CFD studies, and there are currently several numerical approaches that<br />

can be used for this purpose [1]. The objective of a project recently performed<br />

at <strong>Fluent</strong> France and <strong>Fluent</strong> India was to compare these different<br />

approaches in terms of quantitative results and CPU time.<br />

PROCESS INDUSTRIES<br />

A cylindrical, flat bottom, stirred tank with 4 baffles and a non-standard, Rushtontype<br />

impeller with 4 radial blades was used for the simulations. Owing to the<br />

rotational symmetry of the geometry, a 90° sector of the tank was modeled.<br />

The different approaches studied made use of either the multiple reference<br />

frames (MRF) or sliding mesh (SM) model. For the MRF runs, a steady-state<br />

flow field was first computed and a transient calculation of a tracer species<br />

was performed using the “frozen” flow field. The first such blending calculation<br />

was done using the default settings, including the use of relative velocities.<br />

A second blending calculation was performed with absolute velocities.<br />

For this case, the paddles on the impeller were changed to interior zones,<br />

since this setting is more appropriate for walls moving normal to the fluid<br />

when absolute velocities are used. The walls exerting shear on the fluid (the<br />

impeller disk) were not changed. For the transient sliding mesh calculations,<br />

the flow field and tracking of the tracer species were computed simultaneously.<br />

Two solver options were used: the iterative time advancement (ITA)<br />

scheme and the non-iterative time advancement (NITA) scheme, which was<br />

introduced in FLUENT 6.2. The sliding mesh results are the most rigorous, so<br />

they served as a standard for comparison with the other methods. A correlation<br />

was also used to compare the results.<br />

Pathlines illustrate the two circulation loops<br />

in the vessel<br />

A steady state, MRF solution for the flow was first obtained. On this frozen<br />

flow field solution, the transient species calculation was performed after initializing<br />

a volume of tracer in the upper part of the vessel. The MRF flow field<br />

was also used as the starting point for the unsteady sliding mesh calculation.<br />

Before introducing the species, the sliding mesh calculation was done for a<br />

few cycles until periodic behavior was obtained. At this point, the species was<br />

introduced in the same location used for the MRF calculations, and the transient<br />

calculation of the flow field and species was performed.<br />

Results were analyzed globally in terms of the blending time, t 99 , or time<br />

required to achieve 99% uniformity in the tank. A published correlation for<br />

t 99 for a six-bladed Rushton turbine was used for comparison [2]. The blend<br />

time predicted by the Rushton correlation was found to be longer that that<br />

predicted by the CFD methods, a result that is due to the fact that the paddles<br />

on the impeller used have more surface area than a standard Rushton.<br />

In addition, the tracer species introduced from one location in the 90° model<br />

was equivalent to it being introduced from four locations in a 360° vessel.<br />

The tracer species initial locations at t=0<br />

<strong>Fluent</strong> News · Spring 2006<br />

S5


MIXING<br />

PROCESS INDUSTRIES<br />

The tracer distribution after 1.5 (left) and 3.0<br />

(right) seconds for the MRF simulation<br />

mass fraction<br />

mass fraction<br />

6E-03<br />

5E-03<br />

4E-03<br />

3E-03<br />

2E-03<br />

1E-03<br />

SM + ITA<br />

SM + NITA<br />

MRF – ABS<br />

MRF – REL<br />

0E-03<br />

0 2 4 6 8 10 12 14<br />

time (s)<br />

Mass fraction of the tracer species computed by the<br />

SM and MRF methods, recorded at points near the<br />

top of the vessel (top) and near the impeller<br />

(bottom)<br />

6E-03<br />

5E-03<br />

4E-03<br />

3E-03<br />

2E-03<br />

SM + ITA<br />

SM + NITA<br />

MRF – ABS<br />

MRF – REL<br />

The correlation assumes that the species is introduced from one location, usually<br />

at the surface of the liquid. Comparison of the CFD methods revealed that the MRF<br />

model predicted longer mixing times than the sliding mesh model. The two<br />

MRF approaches predicted similar blend times, as did the two SM approaches.<br />

Correlation MRF with MRF with Sliding Sliding<br />

for 6-bladed relative absolute Mesh Mesh<br />

Rushton velocities velocities with ITA with NITA<br />

turbine (default) scheme scheme<br />

Blending time(s): t 99 19.0 (+/- 6) 15.8 (+/- 4) 15.3 (+/- 4) 14.8 (+/- 3) 14.9 (+/- 3)<br />

The various approaches were also compared by monitoring the CPU time required<br />

to reach t 99 . It was found that the sliding mesh model using the new NITA scheme<br />

takes about 4 times less CPU time than the sliding mesh model using the ITA<br />

scheme. The steady-state MRF solutions are much quicker, requiring about 2.5<br />

times less CPU time than the sliding mesh model with the NITA scheme.<br />

The results were also analyzed locally by monitoring the mass fraction of the tracer<br />

species at several sensor locations. Two locations in particular are of interest: near<br />

the tip of the blade and near the top of the tank. Near the impeller, the<br />

sliding mesh case picked up the blade passing frequency, and in addition, a low<br />

frequency oscillation of greater amplitude that was also observed for the MRF<br />

approaches. This low frequency oscillation was also picked up at the top of the<br />

tank by the sliding mesh and MRF approaches, although the blade passing<br />

frequency was not (for the former). The low frequency oscillation is indicative of a<br />

macroscopic exchange of the tracer material between the upper and lower regions<br />

of the vessel. This type of result is not uncommon when a radial impeller is used<br />

and distinct circulation patterns develop above and below the impeller.<br />

Overall, the MRF model is an economic approach for delivering sufficient blending<br />

time information. If, however, an accurate representation of the transient dispersion<br />

of a tracer is required – for reacting flows, for example – the sliding mesh approach<br />

is necessary, and the NITA scheme in FLUENT now makes sliding mesh calculations<br />

much more affordable. <br />

References:<br />

1 Marshall, E.M. and Bakker, A.: Computational Fluid Mixing. In: The Handbook of Industrial<br />

Mixing, Ch. 5. Paul, E., Atiemo-Obeng, V. and Kresta, S., Editors, John Wiley, 2004.<br />

2 Fasano, J.B., Bakker, A. and Penney, W.R.: Advanced Impeller Geometry Boosts Liquid<br />

Agitation. Chemical Engineering, August 1994.<br />

1E-03<br />

0E-03<br />

0 2 4 6 8 10 12 14<br />

time (s)<br />

Iso-surfaces of tracer mass<br />

fraction show the polarized<br />

species distribution in the<br />

vessel, even after the liquids<br />

are well mixed<br />

S6 <strong>Fluent</strong> News · Spring 2006


MICROREACTORS<br />

Liquid Mixing<br />

in Microreactors<br />

By M. Hoffmann, M. Schlueter, N. Raebiger; University of Bremen,<br />

Institute of Environmental Process Engineering (IUV), Bremen, Germany<br />

MICROREACTORS ARE BASIC components<br />

of microfluidic systems for chemical and biochemical<br />

applications and are an important area<br />

of research in fields such as analytical chemistry,<br />

chemical engineering, and life science. They are<br />

well suited for the controlled mixing of reactants<br />

and show great potential for optimizing conventional<br />

processes. The large area-to-volume ratio<br />

of microreactors allows for a higher yield and<br />

selectivity than conventionally designed processes.<br />

To take advantage of the full potential of this<br />

ambitious technology, a fundamental understanding<br />

of the transport processes on the relevant<br />

time and length scales is necessary. One<br />

approach is through the application of CFD. An<br />

example that is currently undergoing study using<br />

numerical simulation is that of a T-shaped<br />

micromixer with a rectangular cross section.<br />

High resolution CFD simulations have been<br />

performed using FLUENT 6.1 at the Institute of<br />

Chemical Engineering at the University of<br />

Paderborn, Germany [1]. The computational<br />

domain is meshed by a block structured grid with<br />

about 970,000 cubic cells. A more refined computation<br />

is also being performed on a grid of<br />

about 9 million cells. For the entrance region in<br />

the mixing channel, inside the mixing zone, the<br />

grids have a spatial resolution of 2.5 and 1µm<br />

respectively. The velocity field is given by the stationary<br />

solution of the Navier-Stokes equations<br />

using a velocity profile of fully developed, rectangular<br />

cross section duct flow at both inlets. At the<br />

outlet, the pressure is set to a reference value. All<br />

channels are closed at the top and a no-slip<br />

boundary condition is applied to all walls. The<br />

properties of water at 20°C are used and a tracer<br />

is added at one of the inlets. The species transport<br />

model is used to track the motion of the<br />

tracer through the micromixer.<br />

In addition to using a commercial CFD program<br />

for numerical flow visualization, microscale fluid<br />

flow visualization is an important tool for acquiring<br />

localized flow information within these<br />

devices. By means of two non-invasive measurement<br />

techniques, micro-laser induced fluorescence<br />

(micro-LIF) and micro-particle image<br />

velocimetry (micro-PIV), the concentration and<br />

velocity fields have been measured.<br />

The tracer distribution at different cross sections<br />

enables a quantitative analysis of the mixer to be<br />

performed. To construct the tracer distribution, a<br />

confocal laser scanning microscope (CLSM) is<br />

used. The CLSM can produce image slices with a<br />

minimum thickness of 2 microns. By combining a<br />

series of these slices, a three-dimensional rendering<br />

of the mixing quality in the mixing channel<br />

is possible. The mixing quality can then be quantified<br />

using Danckwerts’ intensity of segregation<br />

[1]. Other observations made in the CFD analysis<br />

and experimental tests include a secondary flow<br />

in the form of a vortex pair in the entrance<br />

region, which results from instabilities caused by<br />

centrifugal forces.<br />

The results are fundamental for improving the<br />

design rules for static micromixers and have<br />

enabled a validation of the CFD results. <br />

Acknowledgements:<br />

The authors gratefully acknowledge the CFD work<br />

done by our collaborators, Dieter Bothe, Center of<br />

Computational Engineering Science, RWTH Aachen,<br />

Germany and Hans-Joachim Warnecke, Institute of<br />

Chemical Engineering, University of Paderborn,<br />

Germany.<br />

References:<br />

1 Bothe, D., Stemich, C. and Warnecke, H.-J.: Fluid<br />

Mixing in a T-shaped Micro-mixer. Chem. Eng. Sci.<br />

61, 2950-2958, 2006.<br />

Visit us at<br />

ACHEMA 2006:<br />

Hall 1.2, Stand E9-F10<br />

Tracer profile in the y-z plane (top) on a crosssectional<br />

slice through the mixing channel<br />

(bottom) at x=100 µm computed using CFD; the<br />

Reynolds number for the mixing channel is 186<br />

(corresponding to an inlet velocity of 1.4 m/s for<br />

each inlet)<br />

Courtesy of the Institute of Chemical Engineering, University of Paderborn<br />

Velocity field along a mixing channel with a width<br />

of 400 µm and height of 200 µm using micro-PIV;<br />

the measurement plane (x-y plane) is at z=143 µm,<br />

and flow is from left to right<br />

Concentration field in the mixing channel of a<br />

T-shaped micromixer with a width of 400 µm and<br />

height of 200 µm, constructed using image slices<br />

from a confocal laser scanning microscope and<br />

micro-LIF in the y-z plane, with an x range of 270<br />

to 782 µm; the Reynolds number of mixing<br />

channel is 207 and the flow is from back to front<br />

PROCESS INDUSTRIES<br />

<strong>Fluent</strong> News · Spring 2006<br />

S7


THERMAL RUNAWAY<br />

PROCESS INDUSTRIES<br />

Preventing<br />

Runaway<br />

Reaction<br />

Accidents<br />

By D. Dakshinamoorthy and J.F. Louvar, Chemical Engineering Department,<br />

Wayne State University, Detroit, Michigan, USA<br />

RUNAWAY REACTIONS ARE AN ONGOING PROBLEM in the chemical<br />

industry, where they account for 26% of major accidents. Runaway<br />

reactions generate a sudden excess amount of heat, which can lead to<br />

an explosion. They can be stopped in two ways: by the addition of cold<br />

diluents and by the addition of an inhibitor, a chemical that acts to suppress<br />

the runaway reaction. The technology that involves the use of<br />

inhibitors is called shortstopping. Power failures are one of the main<br />

reasons for runaways, and after a power failure, the process of adding<br />

an inhibiting agent and mixing it with the reactor contents becomes a<br />

major problem in the shortstopping process. Jets or impellers, driven by<br />

a small generator, can be used for mixing under such circumstances.<br />

From a design standpoint, jet mixing is one of the simplest methods to<br />

achieve mixing. In jet mixing, a part of the liquid in the tank is drawn<br />

out through a pump and returned to the tank as a high-velocity jet<br />

through a nozzle, resulting in fluid mixing. In a recent project [1], CFD<br />

was used to compare the efficiency of jet mixers with impeller stirred<br />

vessels in shortstopping runaway reactions. On the basis of equal power<br />

consumption, this comparative study showed that jet mixers were ineffective<br />

when used for shortstopping, unless certain factors could be<br />

optimized. Due to the hazardous nature of runaway reactions, these<br />

factors cannot be determined with lab scale or pilot plant scale experiments,<br />

but CFD can be used to carry out virtual experiments instead.<br />

Using FLUENT, mixing with a jet mixer was first investigated for different<br />

nozzle diameters and angles of injection. To account for the external<br />

circulating pump, user-defined functions (UDFs) were used to<br />

ensure that the mass fractions of all species entering the vessel through<br />

the inlet were equal to those leaving through the outlet at each<br />

timestep. Before simulating the reacting flow of the shortstopping<br />

process, the flow model alone was validated. Since overall mixing controls<br />

the process, the predicted mixing times were compared with the<br />

available experimental correlations [2, 3] and found to be in excellent<br />

agreement.<br />

Solution domain and pathlines for the jet<br />

mixer studied<br />

The converged flow results of the best jet configuration were then used<br />

for subsequent simulations of the runaway and inhibition reactions.<br />

Laminar volumetric reactions were modeled using UDFs. Guibert et al. [4]<br />

documented the kinetics of the considered runaway reaction. Before analyzing<br />

a specific runaway scenario, reactor conditions were varied and<br />

runaway behavior was studied. From all of the cases considered, the one<br />

with the fastest temperature increase was selected for further study, since<br />

it would be the most difficult to shortstop. For this scenario, as soon<br />

as the reactor temperature reached 450K, the inhibitor was added and<br />

the species transport equations for the inhibition reaction were solved<br />

simultaneously with the runaway reaction.<br />

The results were used to identify the major and minor factors that contribute<br />

to effective shortstopping when using a jet mixer. These factors<br />

include the location for adding the inhibitor; the amount of the<br />

inhibitor; the rate of the inhibition reaction; the power input; the use of<br />

a cold diluent; and the use of multiple nozzles. The temperature distribution<br />

in the reactor after the shortstopping process and the decrease<br />

in the average reactor temperature were used to assess the importance<br />

of each factor.<br />

S8 <strong>Fluent</strong> News · Spring 2006<br />

Grid near the inlet<br />

For example, the temperature distribution on the mid-plane and an<br />

iso-surface of temperature equal to 500K (temperatures over 500K were<br />

considered hazardous) for two power levels was used to illustrate the<br />

importance of power input. The hotspots, or high temperature regions<br />

over 500K, decrease in size as the power is increased, which indicates<br />

improved shortstopping. In addition to the power input, the decrease in


THERMAL RUNAWAY<br />

mixing time (sec)<br />

10 3<br />

10 2<br />

10 1 Jet Reynolds number<br />

10 2 10 3<br />

10 4<br />

experiment<br />

FLUENT<br />

Comparison of the predicted mixing time and experimental data for a<br />

range of jet Reynolds numbers [2]<br />

the final reactor temperature for the other factors was computed as well. When<br />

compared, the major factors that were found to contribute most to effective<br />

shortstopping were the use of cold diluents and the use of multiple nozzles.<br />

The factors found to be of less importance were the rate of the inhibition reaction,<br />

the location of the inhibitor injection, and the amount of inhibitor. The<br />

study clearly demonstrated the value of using CFD simulations in situations that<br />

are experimentally prohibitive. <br />

Kinetic Details of<br />

Runaway and<br />

Inhibition<br />

Runaway Reaction<br />

Guibert et al. [4] studied the kinetics of the propylene oxide polymerization<br />

reaction. The rate of the reaction is defined as a function of temperature,<br />

propylene oxide concentration, and catalyst concentration.<br />

Monomer polymerizes to polymer in the presence of a basic catalyst. The<br />

kinetic expressions for the monomer concentration and temperature are<br />

as follows:<br />

1.<br />

2.<br />

PROCESS INDUSTRIES<br />

Acknowledgements:<br />

The authors acknowledge NSF for their support of this particular project, and Dr. Vivek.<br />

V. Ranade and his student Dr. Avinash. R. Khopkar of IFMG, National Chemical<br />

Laboratories, Pune, India, for their technical contributions and inspiration.<br />

References:<br />

1 Dakshinamoorthy, D., Khopkar, A.R., Louvar, J.F. and Ranade. V.V.: CFD Simulations<br />

of Shortstopping Runaway Reactions in Vessels Agitated with Impellers and Jets.<br />

Accepted in the Journal of Loss Prevention in the Process Industries, 2005.<br />

2 Lane, A.G.C. and Rice, P.: An Investigation of Liquid Jet Mixing Employing an<br />

Inclined Side Entry Jet. Transactions of Institute of Chemical Engineers, 60, 171-176,<br />

1982a.<br />

3 Lane, A.G.C. and Rice, P.: Comparative Assessment of the Performance of Three<br />

Designs for Liquid Jet Mixing. Industrial & Engineering Chemistry Process Design<br />

and Development, 21, 650-653, 1982b.<br />

4 Guibert, M.R., Plank, A.C. and Gerhard, R.E.: Kinetics of the Propylene Oxide –<br />

Oxypropylated Glycerol Reaction. Industrial & Engineering Chemistry Process<br />

Design and Development, 10 (4), 497-500, 1971.<br />

Inhibition Reaction<br />

The added inhibitor neutralizes the basic catalyst. The kinetic expression<br />

for the inhibition reaction is as follows:<br />

3.<br />

Kinetic Data for<br />

Inhibition and<br />

Runaway Reactions<br />

Activation Energy<br />

Heat of Reaction<br />

Pre-Exponential Factor<br />

Specific Heat Capacity<br />

Pre-Exponential Factor<br />

Gas Constant<br />

E = 6.96E+08 J/kgmol<br />

H = -1.63E+06 J/kg(monomer)<br />

k o = 9.5E+11 kg(total)/(kg(cat)-hr)<br />

c p = 2930 J/kg(total)-K<br />

k 1 = 9.5E+11 kg(total)/(kg(inh)-hr)<br />

R = 8314.34 J/kgmol-K<br />

< 450K > 550K<br />

Contours of temperature on the mid-plane and an iso-surface of tempetature at<br />

500k illustrate the temperature distribution after shortstopping with less (left)<br />

and more (right) power input<br />

<strong>Fluent</strong> News · Spring 2006<br />

S9


EMISSIONS<br />

PROCESS INDUSTRIES<br />

Scrubbers for Flue Gas Cleanup<br />

By Christoph Hochenauer, Austrian Energy & Environment AG, Raaba/Graz, Austria,<br />

Martin Demuth, Institute for Process Technology & Industrial Environmental Protection,<br />

University of Leoben, Austria, and Wolfgang Timm, <strong>Fluent</strong> Germany<br />

The mesh structure<br />

WITH TODAY’S INCREASINGLY RESTRICTIVE environmental<br />

regulations, there is an ongoing need for state-of-theart<br />

modeling of pollutant formation and removal in industrial<br />

processing units. A particularly important issue is the sulphur<br />

dioxide (SO 2 ) content in the flue gases of power plants. A<br />

typical approach for removing this hazardous component is<br />

through absorption by limestone slurry droplets in so-called<br />

flue gas scrubbers. Scrubbers consist of large towers in which<br />

up to several thousands of nozzles provide a homogenous<br />

distribution of droplets within the gas flow. This leads to<br />

the highest possible absorption rate for a given slurry mass<br />

flow. To optimize the nozzle configurations for scrubbers,<br />

CFD simulation is one of the best tools to use.<br />

Mesh in the spray bank zone<br />

The absorption of SO 2 is a complex process with several<br />

mechanisms to be considered. The continuous phase (gas) and<br />

discrete phase (droplets) have to be calculated as well as the<br />

interaction between them. The momentum and temperature<br />

of the droplets affect the gas flow and vice versa.<br />

Additionally, mass transfer between the droplet and gas phases<br />

has to be taken into account. Depending on the difference<br />

in the partial pressure of water vapor in the two phases, there<br />

might be evaporation from the droplets or condensation onto<br />

the droplet surfaces. Rapid changes in temperature cause both<br />

evaporation and condensation to have a significant influence<br />

on the SO 2 absorption rate, which itself is governed by a complex<br />

ion chemistry mechanism within the droplets. The heat<br />

and mass transfer rates are dependent on the droplet surface<br />

area, which changes during evaporation or condensation, or<br />

whenever there is splashing at the walls. To account for the<br />

latter, an appropriate wall interaction model also has to be part<br />

of the CFD simulation. With FLUENT’s user-defined function<br />

(UDF) capability, a set of models addressing these issues has<br />

been implemented, and a test case with typical scrubber<br />

conditions has been created to illustrate the capabilities.<br />

A cylindrical scrubber with two sets of eight radial spray bars<br />

was considered. The grid was built using the polyhedral mesh<br />

S10 <strong>Fluent</strong> News · Spring 2006


EMISSIONS<br />

capability of FLUENT 6.3. The original mesh consisted of a total of 1.25 million<br />

cells, which were reduced to 250,000 cells after the tetrahedral elements<br />

were converted to polyhedra. Using sizing functions, a sufficient grid resolution<br />

in the near wall regions of the spray bank was ensured. A flue gas mass<br />

flow rate of 65kg/s was set at an inlet at the bottom of the unit with SO 2 and<br />

water vapor mass fractions of 0.0024 and 0.12 respectively, and the remainder<br />

consisting of air. Sixty-six full cone nozzles were distributed in a circular<br />

pattern on the sixteen radial pipes for a total slurry mass flow rate of 330kg/s.<br />

Both the gas flow and the droplets started with an initial temperature of<br />

316K. For the droplets, the dispersed phase model (DPM) was used with a<br />

large number of trajectories, whose interaction with the gas phase was evaluated<br />

separately. The RNG k-ε model and the stochastic tracking method<br />

were used to incorporate the effects of turbulent fluctuations on both<br />

phases in the simulation.<br />

In a surrounding air flow at ambient temperature, a water droplet is subject<br />

to a coupled heat and mass transfer process. With increasing velocity<br />

difference between the two phases the transfer is enhanced and can be<br />

calculated with the Ranz-Marshall law, which can be used for SO 2 absorption<br />

as well as vaporization and condensation [1]. Since the latent heat of water<br />

Contours of water vapor content (left), SO 2 mass fraction (middle), and vertical<br />

velocity (right) on four slices within the scrubber; the flow in the unit is from<br />

bottom to top<br />

is quite high, the droplet temperature change along its trajectory cannot be<br />

neglected, since a temperature change of 10K can significantly affect the SO 2<br />

partial pressure at the droplet surface.<br />

When droplets hit the wall there are several possibilities of what can occur.<br />

Under certain conditions a droplet can become part of the liquid film on the<br />

wall and in other instances there may be splashing and breakup of the<br />

droplets. The latter results in an increased number of droplets with a larger<br />

total surface area, a change that again significantly affects the momentum,<br />

heat and mass transfer rates. For this application, it was assumed that the wall<br />

interaction behavior was governed by dimensionless Weber and Reynolds<br />

numbers [2] and a model was developed accordingly.<br />

PROCESS INDUSTRIES<br />

The removal of SO 2 in the droplets is governed by a reaction with limestone,<br />

which can be described by the net reaction:<br />

Droplet temperature evolution along some of the droplet trajectories<br />

Additionally there are several dissolution reactions, since the formation of sulphurous<br />

and carbonic acid results in an ion chemistry mechanism that<br />

depends on the pH-value and the limestone and SO 2 concentrations of the<br />

droplets [3]. This was implemented by means of an externally evaluated<br />

lookup table that was coupled with the discrete phase using FLUENT’s particle<br />

scalar UDF, which allows additional droplet properties to be defined.<br />

Once the total sulphur content in a droplet has been evaluated, the new concentration<br />

of each ion is obtained from the table and stored in the respective<br />

DPM scalar. With this information the calculation of the SO 2 partial pressure<br />

at the droplet-gas interface and the respective sink terms is possible to ensure<br />

SO 2 mass conservation in both phases of the domain.<br />

Droplet splashing and breakup at the walls; the droplet trajectories are colored by<br />

particle diameter, so the red trajectories occur before splashing and breakup occur<br />

The results obtained from the simulation are in very good agreement with<br />

real-life scrubbers. Austrian Energy & Environment AG now includes FLUENT<br />

with these special models in their scrubber design workflow. The polyhedral<br />

cell technology has proved to be a significant improvement in terms of meshing<br />

flexibility and simulation time, enabling engineers to perform a complete<br />

scrubber simulation within a few days. The reliability of the results has meant<br />

that they need to rely less on field tests as well. <br />

References:<br />

1 Aguayo, P. and Weiss, C.: Enthalpy Two Way Coupling for Near Vapor Saturated<br />

Polydispersed Spray Flows. 5th International Conference on Multiphase Flow,<br />

Yokohama, May 30 - June 4, 2004.<br />

2 Mundo, C. and Tropea, C.: Numerical and Experimental Investigation of Spray<br />

Characteristics in the Vicinity of a Rigid Wall. Experimental Thermal and Fluid<br />

Science, 15:228-237, 1997.<br />

3 Weiss, C., Maier, H. and Bärnthaler, K.: Modelling of the Mass Transfer Processes in<br />

the Chemisorption of Flue-Gas Components by Sprays. Proceedings of the ILASS-<br />

Europe 2001; Zürich, 2-6 September 2001.<br />

<strong>Fluent</strong> News · Spring 2006<br />

S11


FURNACES<br />

PROCESS INDUSTRIES<br />

Ultra-low NO x Burners<br />

By Robert J. Gartside and Peter R. Ponzi, ABB Lummus Global, Bloomfield, New Jersey, USA, and David G. Schowalter, <strong>Fluent</strong> Inc.<br />

wall burners<br />

hearth burners<br />

convection<br />

section<br />

The geometry of the ethylene cracking furnace<br />

cracking coils<br />

Get<br />

ENGINEERS AND MANAGERS involved in chemical and<br />

hydrocarbon processing are keenly aware of equipment efficiency<br />

and plant productivity, and the ongoing need for improvement<br />

in these areas. In ethylene production furnaces, light hydrocarbons<br />

are cracked in tubes that are suspended in a combustion<br />

chamber. The tubes have very short residence times. The critical<br />

parameters in the operation of this furnace are the transfer of the<br />

heat of cracking to the tube side hydrocarbons, control of the<br />

tube metal temperatures for prolonged run length, and the<br />

reduction of pollutants such as NO x to meet regulatory statutes.<br />

In the radiant section of a typical furnace, combustion heat is<br />

provided by burning fuel in hearth burners (located on the<br />

floor of the furnace and firing vertically) and in wall burners<br />

(positioned along the wall and firing radially along the wall).<br />

Temperature contours in the original burner<br />

configuration (left) and the first ultra-low NO x<br />

design (right)<br />

The John Zink Company has developed an ultra-low NO x burner<br />

(20 to 30 vppm) in which a portion of the fuel is premixed<br />

with all of the air in both the hearth and wall burners. The<br />

remaining fuel needed to achieve the firing rate is introduced in<br />

a staged manner to control the temperature of the combustion<br />

gases and thereby minimize the production of NO x . This is done<br />

through staged ports located in front of the hearth tile at the<br />

furnace floor. When the introduction of this burner technology<br />

was considered for a client’s ethylene production furnace, it was<br />

suspected that, when compared to the original conventional<br />

burners that generally produce straight vertical flames attached<br />

to the furnace wall, these lean premixed burners might produce<br />

shorter flames that could “roll over” at the bottom of the firebox<br />

and impinge on the tubes. The negative impact of rollover would<br />

be the creation of hot spots on the process tubes and subsequent<br />

increase in coking within the tubes and shortening of tube life.<br />

To investigate this possibility, a realistic CFD study was conducted<br />

to compare the performance of the low NO x and original<br />

burners. The detailed CFD models – one for each type of burner<br />

– included locally refined meshes around the burner inlet ports<br />

and process tubes. Combustion was simulated on the firebox<br />

side to provide the heat generation. A reacting model for the<br />

hydrocarbon cracking that occurs inside the tubes was also<br />

included, as were the effects of turbulence and radiation. Overall,<br />

there was sufficient detail to accurately reflect the heat absorption<br />

and thus provide realistic tube heat fluxes and coil metal<br />

temperatures. The models contained between two and five<br />

Temperatures on three cross-sectional planes in the<br />

original burner configuration (left) and the improved<br />

ultra-low NO x design (right)<br />

S12 <strong>Fluent</strong> News · Spring 2006


MULTIPHASE<br />

Cracking<br />

million computational cells depending on the<br />

type of burner modeled, and they required up<br />

to a week’s worth of CPU time on a high speed<br />

computer cluster.<br />

The initial simulations compared the heater with<br />

the original burners to the heater with the new<br />

burners where the new burners were simply<br />

located at the same position as the original burners.<br />

The results of this study indicated that, while<br />

the original burner design provided straight<br />

flames, the new low NO x configuration resulted<br />

in flames that impinged on the tube surfaces.<br />

Based on these results, an optimization study<br />

using CFD was done to develop design modifications<br />

that would allow the ultra-low NO x burner<br />

to perform acceptably in the heater. These modifications<br />

included redirecting the angle of the<br />

fuel that was being injected through the ports in<br />

front of the new burners and relocating the burners<br />

on the furnace floor to provide lateral spacing<br />

into which the combusting gas could expand,<br />

thereby reducing the tendency for the flames to<br />

roll into the tubes. The modifications served to<br />

straighten the flames, making them more in line<br />

with those of the original burners. While the new<br />

design produces more diffuse flames than the<br />

original burner design, it is much improved from<br />

the first ultra-low NO x design.<br />

The new burner configuration, optimized by<br />

CFD, has been installed in the client’s furnace.<br />

Flame quality and run length have been excellent<br />

and low NO x operation has been experienced.<br />

The start-up was smooth and downtime was<br />

minimal. <br />

Acknowledgement:<br />

The authors wish to acknowledge the cooperation<br />

of the John Zink Company in the low NO x burner<br />

modification CFD studies.<br />

Controlling Droplet<br />

Size Distribution in<br />

Emulsions<br />

By L. Srinivasa Mohan, <strong>Fluent</strong> India, and Ahmad Haidari and Aniruddha Mukopadhyay, <strong>Fluent</strong> Inc.<br />

Emulsions are an important class of materials produced<br />

and handled by the chemical, food, pharmaceutical,<br />

and cosmetic industries. They consist<br />

of two immiscible liquids, one dispersed in the<br />

other. The properties of an emulsion are based<br />

on the droplet size distribution (DSD) of the<br />

dispersed phase. Because they are often thermodynamically<br />

metastable, there is a persistent<br />

threat that the texture of the emulsion will be<br />

altered during the course of preparation or packaging,<br />

or during the subsequent shelf life. Many<br />

processes over widely varying length scales could<br />

cause the DSD to change, and it is important that<br />

they be well understood so that the emulsion<br />

quality can be maintained.<br />

For an ongoing project at <strong>Fluent</strong>, several aspects<br />

of droplet behavior have been studied using the<br />

volume of fluid (VOF) model. Some of the results<br />

have been compared to experiments that were<br />

carried out on microfluidic devices, where a precise<br />

droplet size distribution could be generated.<br />

For example, the generation of droplets at a<br />

T-junction using two streams of immiscible liquids<br />

has been simulated. Droplets of uniform size<br />

were rapidly and reproducibly produced at the<br />

junction as a result of the surface tension and the<br />

shearing motion of the fluid in the main channel,<br />

in agreement with measurements.<br />

In another study, the geometrically mediated<br />

breakup of droplets in a microfluidic device [2]<br />

was simulated. By changing the length of the<br />

arms of the T-junctions, the droplet can be split<br />

into daughter droplets of unequal size. By using a<br />

network of asymmetric T-junctions, emulsions of<br />

a given DSD can be produced. Both 2D and 3D<br />

simulations matched the qualitative and quantitative<br />

aspects of the experiments, such as the size<br />

of the daughter droplets for a given T-junction<br />

and the critical parameters required for droplet<br />

breakup as a function of capillary number. <br />

References<br />

1 Nisisako, T., Torri, T. and Higuchi, T.: Lab Chip. Vol<br />

2, no 1. pp 24-26, 2002.<br />

2 Link, D.R., Anna, S.L., Weitz, D.A. and Stone, H.A.:<br />

Phy. Rev. Lett. Vol. 92, pp 1178-1180, 2004.<br />

1E-07<br />

8E-08<br />

6E-08<br />

4E-08<br />

2E-08<br />

volume of droplet (m 3 )1.2E-7<br />

0<br />

0.0<br />

continuous phase flow<br />

dispersed phase flow<br />

Schematic of the experimental setup [1]<br />

0.1 0.2 0.3 0.4 0.5 0.6 0.7<br />

time (s)<br />

Volume of droplets produced at the T as a<br />

function of time<br />

Droplet breakup at a symmetric (left) and<br />

an asymmetric (right) T-junction<br />

PROCESS INDUSTRIES<br />

<strong>Fluent</strong> News · Spring 2006<br />

S13


MULTIPHASE<br />

PROCESS INDUSTRIES<br />

Understanding<br />

Fluid-Bed<br />

Coating<br />

Rod Ray, Lisa Graham, Rick Falk, Josh Shockey, and Leah Appel,<br />

Bend Research Inc., Bend, Oregon, USA, and Kumar Dhanasekharan,<br />

<strong>Fluent</strong> Inc. and L. Srinivasa Mohan, <strong>Fluent</strong> India<br />

DURING THE PAST DECADE, the use of fluidized bed systems for<br />

coating has developed to a point where this equipment is now used<br />

for applications in industries ranging from chemical to pharmaceutical to<br />

agriculture. The appeal of fluidized beds for coating is due to the high energy<br />

and mass transfer rates that are available in such a system. There are<br />

three types of fluidized beds that are typically used for this purpose:<br />

top-spray, tangential-spray, and bottom-spray. In a recent joint project<br />

involving engineers at Bend Research and <strong>Fluent</strong>, a fluidized bed with a<br />

bottom-spray of the Wurster type was studied.<br />

Fluid bed coaters are used to obtain controlled release or delayed release<br />

particles in pharmaceutical and agricultural applications. For these types of<br />

coatings, it is important to have control of both coating quality and coating<br />

uniformity. Typically, this process is run in batch mode with particles<br />

circulating through the coating zone multiple times in order to achieve the<br />

desired coating level. In this process, the particles are entrained in a high<br />

velocity gas, they pass through the Wurster column where coating is applied<br />

via a two-fluid atomizer, they then dry in the expansion chamber and<br />

fall back down to the bed to repeat the process. FLUENT has been used to<br />

help understand the process, design process conditions that yield products<br />

within an acceptable variation, and understand the effect of various<br />

parameters on product variability and quality.<br />

coating solution<br />

atomization air<br />

fluidization air<br />

Wurster gap<br />

Schematic illustration of a process stream in a<br />

bottom-spray Wurster fluidized bed coater<br />

Accounting for the presence of a secondary phase within CFD software is<br />

well established. Some of the more recent advances enable simulations of<br />

flows with very high solid loading, such as hoppers and chutes; flows with<br />

large particles, such as particle milling machines; gas-solid flows with particle<br />

size variation, such as polymerization reactors; and flows where the effects<br />

of particle shapes and particle mechanics are important. For the Wurster bed<br />

simulation, the Eulerian granular model in FLUENT was used. The results<br />

have led to a better understanding of the uniformity of coating, spray<br />

patterns, and fluidization behavior. Parameters that control the process<br />

efficiency and coating uniformity were studied. These included the<br />

equipment geometry, air flow rate, and other spray-related characteristics.<br />

Of particular importance for the equipment geometry is the Wurster gap, a<br />

small passage at the base of the unit between the inner and outer (annular)<br />

columns. During operation, the particles in the bed are fluidized and<br />

coated primarily in the inner column, where the fluidization velocity is<br />

highest. As a result, a steady fountain of solid particles is produced, and after<br />

reaching the upper region of the device, these particles stream down the<br />

walls of the outer section prior to being entrained through the gap and<br />

re-entering the primary fluidization region.<br />

The results indicated that there is a steady solids flux through the larger gap,<br />

but in the case of the smaller gap, the flow of solids through the gap is<br />

choked periodically and the bed does not maintain a steady fountain. This<br />

creates a non-uniform solids flux in the spray region, which could lead to<br />

problems with coating uniformity. <br />

Contours of particle volume fraction for two different Wurster gap designs:<br />

wide (left) and narrow (right), for a similar air flow rate<br />

S14 <strong>Fluent</strong> News · Spring 2006


PUMPS<br />

Pathlines through one of<br />

Tuchenhagen’s pump designs<br />

Pumping out<br />

New Designs<br />

More Quickly<br />

By Matthias Südel, Tuchenhagen GmbH, Büchen, Germany<br />

TUCHENHAGEN has been producing process components<br />

for the brewery, beverage, dairy and food industries for more<br />

than seventy years. In addition, the valves, pumps, and cleaning<br />

devices they produce are also used in the cosmetics,<br />

healthcare, pharmacy, and fine chemicals segments, where<br />

hygiene and sterility play a major role. The company, which<br />

has been a member of the GEA group since 1999, introduced<br />

the first twin seat valve with a mixproof technology in 1967.<br />

Customers’ expectations about the quality of process technology<br />

used have been increasing ever since. Many products<br />

must meet the stringent requirements outlined in the<br />

European Hygienic Equipment Design Group (EHEDG) and 3A<br />

sanitary standards. All devices must be capable of cleaning<br />

and flushing and must be free from dead space (clearance<br />

volume).<br />

In the past, designers made predictions about the cleaning,<br />

purification, and purging capabilities of the new components<br />

only after expensive technological trials. Today, CFD is capable<br />

of mapping these prototypes, allowing designers to examine<br />

every corner of the valve, pump, or cleaning device.<br />

FLUENT is now used to accelerate process safety and the<br />

durability of Tuchenhagen’s components. The software makes<br />

flow patterns visible and tracks even the smallest of problem<br />

areas in the course of computations. As a result, the development<br />

costs associated with the recent generation of pumps<br />

have decreased by seven percent compared to the conventional<br />

development process. The time required to reach the<br />

mass production phase has been reduced by six to eight<br />

weeks for each new design.<br />

For many years, engineers at Tuchenhagen could design and<br />

build a new pump based solely on their years of experience.<br />

With CFD, the company now has at its disposal a fast and costeffective<br />

method to support this experience with numerical<br />

data. Using this principle, Tuchenhagen developed three new<br />

pumps last year. Their approach is to carry out a computation,<br />

then built the prototype and make measurements.<br />

To improve upon this process, they have defined special<br />

configuration attributes in the CFD software. In two cases, the<br />

pumps were modeled using CFD, then tested at the operating<br />

point, and quickly put into mass production. A typical pump<br />

simulation makes use of a hybrid mesh of about 1 million cells.<br />

The steady-state MRF model is used to simulate the rotation of<br />

the moving parts. The standard k-ε model with enhanced wall<br />

treatment is used for turbulence.<br />

PROCESS INDUSTRIES<br />

A VARIVENT ® mixproof<br />

valve Type R<br />

Velocity magnitude on different<br />

planes of a VARIVENT ® valve<br />

Type R<br />

The numerical simulations have reduced the time required<br />

to go from the first prototype to the actual mass-produced<br />

product. Additionally, material costs for high-grade steel<br />

during the prototyping phase do not need to be considered.<br />

Typically, making the first sample takes up to two weeks, and<br />

measurements take up another two to four weeks depending<br />

on the components. This is partly due to the fact that the<br />

cleanability of the products, which is so important to<br />

Tuchenhagen, must be repeatedly demonstrated in product<br />

trials. While it is not possible to entirely eliminate test runs, the<br />

company can now examine predictions of the wall shear<br />

stress, which is the deciding quantity for cleanability, since it<br />

is related to the ability of the fluid to dislodge dirt particles. <br />

<strong>Fluent</strong> News · Spring 2006<br />

S15


EXTRUSION<br />

PROCESS INDUSTRIES<br />

Polymer Processing<br />

Simulation for Foam Extrusion<br />

By Hirohisa Shiode and Takeharu Isaki, Mitsui Chemicals, Inc., Material Science Laboratory, Chiba, Japan<br />

MITSUI CHEMICALS and Mitsui Chemicals<br />

Group produce a wide variety of chemical products<br />

from catalysts to polymers and fabricated<br />

products like films and fibers. The Computational<br />

Science Department at the Material Science<br />

Laboratory investigates simulation technologies<br />

and applications to research and development in<br />

the areas of chemistry, engineering, polymeric<br />

material design and polymer processing.<br />

Since 1993, POLYFLOW has been used for polymer<br />

processing applications involving various<br />

rheology models, boundary conditions, and free<br />

surface flows, such as extrusions through<br />

complex and multi-layered dies and profile<br />

extrusions. One area of interest is foam extrusion,<br />

which includes gas dissolution into the polymer<br />

melt, viscosity reduction, nucleation with pressure<br />

drop, and bubble growth. The foam properties<br />

are closely related to cell structures, which are<br />

determined by the process conditions. A simulation<br />

method for gas-dissolved polymer flow has<br />

been developed for predicting these cell<br />

structures and the optimal die design. It makes<br />

use of a viscosity model, which is based on free<br />

volume theory [1,2] and which takes into<br />

account the temperature, pressure, and gas<br />

concentration in the melt. The model is<br />

implemented in POLYFLOW by means of userdefined<br />

functions (UDFs), and is used for 3D<br />

simulations of the flow through the die. To<br />

account for the shear-thinning behavior of the<br />

material, the Cross-Carreau model is adopted.<br />

The pressure distribution in a coat hanger die with<br />

a gas concentration of 1%<br />

pressure drop (MPa)<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0.0<br />

The geometry of a coat hanger die<br />

0% gas<br />

1% gas<br />

2% gas<br />

3% gas<br />

0.2 0.4 0.6 0.8 1.0<br />

dimensionsless length<br />

The pressure drop<br />

through a coat<br />

hanger die along the<br />

centerline for gas<br />

concentrations<br />

ranging from 0 to 3%<br />

As an example, the customized model has been<br />

applied to the flow through a coat hanger style<br />

die. The results indicate that the pressure drop<br />

through the die decreases with increasing gas<br />

concentration as a result of the associated drop in<br />

viscosity. This result is consistent with experimental<br />

observations. Further work is now being conducted<br />

to couple the flow simulations with nucleation<br />

and cell growth. <br />

References<br />

1 Shiode, H. and Isaki, T.: Polymer-Supercritical Fluid<br />

Systems and Foams. Dec., Tokyo P-22 188, 2003.<br />

2 Willams, M.L., Landel, R.F. and Ferry, J.D.: J. Am.<br />

Chem. Soc., 77, 3701, 1955.<br />

S16 <strong>Fluent</strong> News · Spring 2006


AEROSPACE<br />

The impact of a rear-mounted nacelle on the top surface<br />

of a wing is illustrated by static pressure contours<br />

Nacelle Impact<br />

on Aircraft Wing & Fuselage<br />

By Zhu Jie, AVIC 1 (Aviation Industries of China) Commercial Aircraft Company, Shanghai, China<br />

THE GENERAL CONFIGURATION and aerodynamic<br />

design integration of the nacelle on a<br />

commercial aircraft have an important influence on<br />

economy, safety, comfort, and performance. It is a<br />

complex yet important task to study the aerodynamic<br />

interference between the wing/body and<br />

engine nacelle and develop techniques for integrating<br />

these components during the design phase.<br />

These considerations have played a role in the<br />

development of the advanced regional jet ARJ21<br />

from China’s AVIC 1 Commercial Aircraft Company<br />

(ACAC). At ACAC, FLUENT was introduced during<br />

the pre-development phase, and it continues to<br />

play a major role in design, engine selection, and to<br />

save time spent on wind tunnel tests.<br />

Before applying the turbulent Navier-Stokes<br />

equations to the complete aircraft, validations were<br />

performed on a simplified geometry. Using a 3D<br />

wing-body model, predictions of pressure and lift<br />

coefficient were compared to experimental data,<br />

and very good agreement was obtained. These<br />

results gave the engineers confidence in CFD, and<br />

its use has expanded to include many components<br />

and aspects of the full-scale aircraft.<br />

As an example, the flow above the joint that is<br />

formed between the wing and body of an aircraft has<br />

been examined. The flow in this region can separate,<br />

and if it is possible to reduce or eliminate the size of<br />

the separation region by design modifications, the<br />

flight performance of the plane can be improved.<br />

Engineers at ACAC have used FLUENT to study this<br />

phenomenon and improve the flow characteristics.<br />

Pressure (Form) Drag Friction Drag Total Drag<br />

Clean wing 0.01547 0.00412 0.01959<br />

Wing with a rear-mounted nacelle 0.01367 0.00397 0.01764<br />

As another example, a rear-mounted nacelle configuration<br />

has been studied, in which the aircraft’s<br />

engines are installed behind the wings. In this configuration,<br />

the position of the nacelle or the design<br />

of the wing can cause the wake of the wing to be<br />

entrained by the engine, compromising its performance.<br />

Using FLUENT, the pressure distribution on<br />

the wing for several different rear-mounted nacelle<br />

positions has been computed for a flight Mach<br />

number of 0.78 at an altitude of 35,000 feet. The<br />

results showed that the lift coefficient of the wing<br />

with this configuration is lower than that of a clean<br />

wing (without the interference of the nacelle). The<br />

drag coefficient is also lower, however, so that the<br />

lift/drag ratio is increased. Furthermore, the interference<br />

caused by rear-mounted nacelles makes the<br />

nose-down pitching moment decrease and it can<br />

reduce the trim drag produced by the elevators. The<br />

overall analysis of drag indicates that the existence<br />

of a rear-mounted nacelle will have little influence<br />

on friction drag, but can have the advantage of<br />

reducing the pressure drag. <br />

The influence of a rear-mounted engine nacelle on components of drag for a case with a wing lift coefficient of 0.46<br />

CP<br />

-1.5<br />

-1.0<br />

-0.5<br />

experiment<br />

FLUENT<br />

Z = 0.75<br />

Z = 0.37<br />

0.0<br />

0.5<br />

Static pressure on the<br />

surface of the ARJ21<br />

0.0<br />

0.2 0.4 0.6 0.8 1.0<br />

x<br />

A comparison of the pressure coefficient computed<br />

by FLUENT and measured in the wind tunnel for<br />

two wing span locations for the simplified wing-body<br />

geometry<br />

Flow separation near the wing-body joint (lower left)<br />

was improved by using CFD to alter the design<br />

(lower right)<br />

<strong>Fluent</strong> News · Spring 2006 19


AEROSPACE<br />

Compressors Benefit from<br />

the<br />

NASA Rotor 37<br />

By Răzvan Mahu and Cristina Oprea, TENSOR, Bucharest, Romania<br />

percent span<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

1.24<br />

experiment<br />

FLUENT<br />

1.28 1.32 1.90 2.10 2.30<br />

total pressure ratio<br />

The total temperature ratio (left) and total<br />

pressure ratio (right) along the wing span,<br />

ranging from the blade root (0%) to the blade<br />

tip (100%), approximately 10cm from the<br />

leading edge at 98.7% of the maximum<br />

(experimentally measured) flow rate (20.93kg/s)<br />

2.2<br />

2.1<br />

2.0<br />

1.9<br />

19.0<br />

experiment<br />

FLUENT<br />

19.5 20.0 20.5 21.0<br />

mass flow (kg/sec)<br />

The predicted compression<br />

ratio is within 1% of the<br />

experimental values<br />

throughout the range of<br />

mass flow rates studied<br />

THE NASA ROTOR 37 is among the most popular rotors that have been manufactured<br />

to date. It was designed and tested in the early 1980s by Reid and<br />

Moore at NASA Glenn Research Center [1]. It consists of 36 blades with multiple<br />

arc profiles, and was designed for compressors and turbines with a compression<br />

ratio of 2.05 at a mass flow rate of 20.19 kg/s.<br />

Steady-state numerical simulations of the NASA Rotor 37 have been performed<br />

using FLUENT for one operating curve. Using a constant rotational speed (17,188<br />

rpm), the simulations covered a range that extends from the maximum flow rate<br />

to the pumping regime. A fully structured mesh of 420,000 cells was built using<br />

GTurbo, <strong>Fluent</strong>’s turbomachinery preprocessor. During the preliminary runs, the<br />

shock wave accuracy was found to strongly depend on the mesh resolution<br />

around the blade. Consequently, most of the cells were concentrated near and<br />

between the blades. A comparison of turbulence models was also conducted<br />

early on, and the most suitable one for this case was found to be the realizable<br />

k-ε model. Since the mean y + value was around 33, non-equilibrium wall<br />

functions were used. An axial flow direction was set at the inlet, and rotationally<br />

periodic lateral boundaries were used. An inlet temperature of 288.16K was<br />

assumed, and the exit pressure was altered for each of seven points on the<br />

operating curve that were simulated.<br />

Using data from published experimental results [2] a comparison of the radial<br />

variation of the pitch-averaged total pressure ratio was made, 10.19 cm from the<br />

blade leading edge. The results indicate that the total pressure is lower at the<br />

blade tip (100% span) resulting in losses due to the tip gap. A comparison of the<br />

total temperature ratio along the blade span, measured at the same location,<br />

shows an increase in the total temperature at the blade tip due to the friction<br />

between the fluid and the shroud, and to the detaching of the boundary layer.<br />

In both cases, the FLUENT results were in very good agreement with the data.<br />

Contours of Mach number at 70% of the blade height<br />

at 98.7% of the maximum experimental flow rate<br />

Predictions of the efficiency of the compression process along the operating<br />

curve were found to be good at low flow rates, but to deviate from the data at<br />

large flow rates. The total temperature ratio along the operating curve was found<br />

to be in good agreement throughout the range, even though the computed<br />

values were higher. A comparison of the compression ratio as a function of the<br />

flow rate was found to be within 1% throughout the range studied.<br />

Contours of Mach number on a plane through the blades illustrate some of the<br />

characteristics of flow in transonic compressors. The detached bow shock wave<br />

at the leading edge of each blade generates a normal shock wave in the flow<br />

channel that is the main compression mechanism for this type of compressor.<br />

There is a low momentum fluid region behind the shock wave, which is the<br />

second main source of total pressure loss and compression efficiency. <br />

References<br />

1 Reid, L. and Moore, D.: Performance of Single-Stage, Axial-Flow Transonic Compressor<br />

With Rotor and Stator Aspect Ratios of 1.19 and 1.26 Respectively, and with Design<br />

Pressure Ratio of 2.05. Tech Rep. TP-1338, NASA.<br />

2 NASA/TM-2003-212457 Report, http://gltrs.grc.nasa.gov<br />

Pathlines colored by relative Mach number at 98.7%<br />

of the maximum experimental flow rate<br />

20 <strong>Fluent</strong> News · Spring 2006


FOOD<br />

Image-based<br />

Easy as Pie<br />

Meshing:<br />

By Emma Johnson, Roderick Ross and Philippe Young, Simpleware Ltd., Exeter, UK; Gavin Tabor, University of Exeter, Exeter, UK<br />

IMAGED-BASED MESHING is opening up exciting new possibilities for the<br />

application of computational continuum mechanics (numerical methods such as CFD<br />

and finite element analysis (FEA)) to problems in biomechanics, biofluid dynamics and<br />

materials characterization. Software solutions that can rapidly generate robust, high<br />

quality meshes from complex 3D image data, as can be obtained from magnetic resonance<br />

imaging (MRI), computed tomography (CT) or ultrasound for example, are<br />

increasingly in demand. Such software improves productivity, leads to significantly<br />

more accurate results and enables engineers to focus on the analysis and generation<br />

of results rather than on the geometry definition and mesh creation.<br />

The University of Exeter and Simpleware Ltd. have collaborated to illustrate the<br />

potential of new meshing techniques by generating a CFD model of a hot mince pie,<br />

cooling in a light breeze. The transient study uses Simpleware software to create a<br />

mesh from medical scan data and FLUENT to analyze the coupled heat transfer and<br />

fluid flow of the system. The example illustrates a number of more general points<br />

pertaining to multi-physics simulations on complex image-based domains. First, the<br />

method allows watertight surface meshes and fully volumetric meshes of complex<br />

structures to be generated. Second, the accuracy of the geometric reconstruction is<br />

only dependent on the image quality, which is a function of the scan resolution, noise,<br />

and contrast between the volumes of interest. In fact, because structures are defined<br />

by iso-surfaces that are interpolated across boundaries, the geometric reconstruction<br />

can be of sub-voxel accuracy. (A voxel is a volume pixel, or smallest volume for which<br />

image data can be stored.) Third, the method allows multiple structures to be<br />

identified and meshed, with perfectly conforming/non-overlapping boundaries, and<br />

with the option of defining contact surfaces between them. Finally material properties<br />

within a given structure can be assigned based on signal strength. In the pie example,<br />

thermal properties could be varied throughout the mesh of the crust based on fat<br />

content, for example, and the different regions could then be modeled as separate<br />

solid zones.<br />

The segmentation of structures (crust, filling, raisins and air) in<br />

ScanIP software<br />

A 60g mince pie was scanned using MRI. Four segmentation masks, representing the<br />

crust, filling, raisins, and air, were generated using Simpleware’s ScanIP segmentation<br />

software. A volumetric mesh was then generated in the +ScanFE mesh generation<br />

module and exported directly to FLUENT. The time required to go from the import of<br />

the 3D data through to the creation of the multi-part input mesh for FLUENT was less<br />

than 10 minutes on a PC.<br />

In FLUENT the physical properties were assigned and the filler temperature was set to<br />

450K. A rectangular space was created around the pie with boundaries 30cm downstream<br />

and 4cm on either side of the pie. Upstream of the pie, an inlet condition was<br />

set with a velocity of 2m/s and temperature of 300K, akin to a short term cooling of<br />

the pie in a breeze. These boundary conditions were used for a transient calculation<br />

during which the filler was allowed to cool naturally for a time of 300s (5 minutes).<br />

The results show features such as the temperature gradients inside the pie and the<br />

recirculation zone downstream of the pie and inlet. <br />

Temperature contours and surfaces inside the pie, and<br />

velocity vectors showing the cooling breeze at the start of<br />

the calculation (top) and after several minutes (bottom)<br />

More.info@<br />

www.simpleware.com<br />

<strong>Fluent</strong> News · Spring 2006 21


FOOD<br />

Looking Inside<br />

Dough Mixers<br />

By Robin Connelly, Departments of Food Science and Biological Systems Engineering, University of<br />

Wisconsin, Madison, Wisconsin, USA, and Jozef Kokini, Department of Food Science, Center for Advanced<br />

Food Technology, Rutgers University, New Brunswick, New Jersey, USA<br />

View of the C.W. Brabender, Inc. Farinograph ® twin<br />

sigma blade mixer<br />

Mixing Index for Newtonian corn syrup for two<br />

mixer positions, where a value of 0 (blue) indicates<br />

pure rotation, 0.5 indicates simple shear, and 1 (red)<br />

indicates pure elongation<br />

22 <strong>Fluent</strong> News · Spring 2006<br />

WHEAT FLOUR DOUGH is a rheologically<br />

complex viscoelastic material, whose unique timedependent<br />

properties are governed by the rate,<br />

amount, and type of deformation applied. The<br />

structure and morphology of dough also depend on<br />

the available moisture and the extent to which the<br />

dough is mixed. As a result, dough is a very dynamic<br />

and unstable material, and dough mixers have<br />

evolved into highly complex geometries that shear,<br />

stretch, and fold. In addition, they often have close<br />

wall clearances to ensure that there are no regions<br />

of ineffective mixing. Changing between mixer<br />

types, especially between batch and continuous<br />

mixers, is difficult because of the very different flow,<br />

shear, extension, and mixing profiles that characterize<br />

each one. In industry, determining mixing times<br />

and designing mixer configurations is largely done<br />

on a trial and error basis.<br />

A more predictive approach is to calculate measures<br />

of mixing effectiveness through the use of mathematical<br />

modeling and numerical simulation. An<br />

ongoing project at the Center for Advanced Food<br />

Technology at Rutgers, The State University of New<br />

Jersey, supported by the National Research Initiative<br />

of the USDA Cooperative State Research, Education<br />

and Extension Service*, is using POLYFLOW to study<br />

the flow and mixing in typical dough mixer geometries<br />

with fluid models that range from simple<br />

Newtonian to non-linear viscoelastic [1,2]. The work<br />

featured here involves the numerical simulation of<br />

the flow and mixing in a fully filled Farinograph ®<br />

mixer using the mesh superposition technique and<br />

particle tracking with a Newtonian fluid model<br />

based on a high viscosity corn syrup [3]. Future<br />

work with this geometry will focus on extending the<br />

simulations to non-Newtonian fluid models. The<br />

Farinograph is a low shear rate batch mixer with<br />

two non-intermeshing, asymmetrical sigma blades,<br />

where the fast (right) blade turns at 93 rpm counterclockwise<br />

and the slow (left) blade turns at 62<br />

rpm clockwise. CAD STEP representations of the<br />

blade geometries were provided by C.W. Brabender<br />

Instruments, Inc., South Hackensack, New Jersey,<br />

producer of the Farinograph Mixer.<br />

Because the blades turn at different speeds, two<br />

revolutions of the slow blade and three revolutions<br />

of the fast blade are required before there is repetition<br />

of the relative blade positions. The left (slow)<br />

blade mesh of 6232 tetrahedral elements and the<br />

right (fast) blade mesh of 6166 elements are superimposed<br />

on the bowl (41860 hexahedral elements)<br />

every 0.027 seconds, giving a total of 72 positions<br />

per blade cycle with 10° between positions for the<br />

slow blade and 15° between positions for the fast<br />

blade. The time marching flow simulation results are<br />

then used to generate particle tracking data for<br />

10,000 massless material points initially randomly<br />

distributed throughout the flow domain, or a set of<br />

1000 massless material points initially randomly<br />

distributed in a 1 cm 3 box. As these abstract points<br />

are tracked throughout the flow domain, the associated<br />

local flow characteristics are recorded, thus<br />

providing a spatial and temporal history of phenomena<br />

such as stretching and deformation. Random<br />

distribution is a requirement of the statistical mixing<br />

measures that are used to evaluate the particle<br />

tracking results.<br />

The simulation results show that the differential in<br />

the speed of the two blades in the Farinograph<br />

causes an exchange of material between the blades<br />

to occur. The primary circulation pattern consists of<br />

material moving from the slow blade up toward the<br />

top of the mixer and over toward the fast blade,<br />

while the fast blade pushes material towards the<br />

slow blade near the bottom of the mixer. A slower<br />

mixing pattern is also observed where material<br />

around the blades moves from the center towards<br />

the walls and then up towards the top and back<br />

down in the center of the mixer. The zone in the


FOOD<br />

center of the mixer between the two blades is<br />

shown to have excellent distributive and dispersive<br />

mixing ability with high shear rates and mixing<br />

index values [4]. The mixing index is a measure of<br />

the type of flow, with values that range from 0 for<br />

pure rotational flow to 0.5 for shear flow to 1 for<br />

pure elongational flow. A high value of the mixing<br />

index combined with a high shear stress and shear<br />

rate indicate an area in the mixer with potentially<br />

good dispersive mixing capability. That region also<br />

has fast distribution throughout both sides of the<br />

lower section of the mixer as shown by material<br />

point clusters that travel through it. In contrast, very<br />

slow mixing is seen in the area away from the region<br />

swept by the blades that is generally not filled<br />

during normal use of this mixer.<br />

The mean of the normalized length of stretch [5]<br />

calculated for material points in the Newtonian fluid<br />

case increased exponentially over time, indicating<br />

effective mixing for the majority of material points.<br />

In the area swept by the blades, the material points<br />

with the highest values of the length of stretch are<br />

generally located near the blade edges or in the area<br />

swept by the blade edges. However, material points<br />

with high length of stretch values are also found<br />

outside these zones in a more random distribution.<br />

The instantaneous efficiency, which can be thought<br />

of as the fraction of energy dissipated locally that is<br />

used to stretch a fluid element at a given instant in<br />

a purely viscous fluid [5], gives a picture of the most<br />

and least effective blade positions for applying energy<br />

to stretch rather than displace material points.<br />

The least effective blade positions are when the flattened<br />

central sections of both blades are horizontal,<br />

while the most effective mixing occurs when the<br />

flattened section of the fast blade is vertical. The<br />

mean time-averaged-efficiency [5] stays above<br />

zero while its standard deviation reduces over time,<br />

indicating that the majority of the points are<br />

continuously experiencing stretching at equivalent<br />

levels over time.<br />

The overall result of this project has been to<br />

demonstrate the effectiveness of numerical simulation<br />

as a means to non-intrusively study the<br />

flow and mixing in a given mixer of materials<br />

with different rheological properties. The results<br />

also have established a basis of comparison<br />

between mixers with very different geometries.<br />

The simulations are providing guidance for the<br />

design of experiments that will be used to<br />

validate the findings of the simulations. Once the<br />

simulations are validated, they will be able to<br />

provide a much higher level of detail than the<br />

experimental results. The insight gained by this<br />

research has already stimulated the development<br />

of ideas by those involved on how the flow and<br />

mixing in a mixer affects the development of<br />

material structure, leading to profitable new lines<br />

of research. <br />

References:<br />

1 Connelly, R.K. and Kokini, J.L.: Analysis of Mixing in a<br />

Model Mixer using 2-D Numerical Simulation of<br />

Differential Viscoelastic Fluids with Particle Tracking.<br />

J. Non-Newt. Fluid Mech., 123:1-17, 2004.<br />

2 Connelly, R.K. and Kokini, J.L.: 2-D Numerical<br />

Simulation of Differential Viscoelastic Fluids in a Single-<br />

Screw Continuous Mixer: Application of Viscoelastic<br />

FEM Methods. Adv. Poly. Tech., 22(1):22-41, 2003.<br />

3 Connelly, R.K. and Kokini, J.L.: Mixing Simulation of a<br />

Viscous Newtonian Liquid in a Twin Sigma Blade Mixer.<br />

AIChE J., In Review, 2006.<br />

4 Yang, H.-H., Wong, T.H. and Manas-Zloczower, I.:<br />

Flow Field Analysis of a Banbury Mixer. In: Mixing and<br />

Compounding of Polymers: Theory and Practice, Ch. 7,<br />

pp. 187-223, Manas-Zloczower, I. and Tadmor, Z.<br />

Editors, Carl Hanser Verlag, New York, 1994.<br />

5 Ottino, J.M.: The Kinematics of Mixing: Stretching,<br />

Chaos and Transport. Cambridge University Press,<br />

1989.<br />

6 Connelly, R.K. and Kokini, J.L.: 3D Numerical<br />

Simulation of the Flow of Viscous Newtonian and Shear<br />

Thinning Fluids in a Twin Sigma Blade Mixer. Adv. Poly.<br />

Tech., In Review, 2006.<br />

* Grant numbers 2001-35503-10127 & 2003-35503-<br />

13907.<br />

Newtonian corn syrup fluid velocity vectors for blade<br />

positions of 180° and 270° on the z=0 (top), x=0<br />

(middle), and y=4.225 cm (bottom) planes; the vectors<br />

are all the same length and colored by the velocity<br />

magnitude<br />

Initial (left) and final (right) positions after 3 blade cycles (6 revolutions of the slow blade and 9 revolutions of the<br />

fast blade) of 1,000 material points in the Newtonian corn syrup, colored according to the length of stretch [6]<br />

For the Newtonian corn syrup, 3D positions of<br />

10,000 initially randomly distributed material points<br />

(top) and after three cycles (bottom) with<br />

concentrations of 1 (red) and 0 (blue) [6]<br />

<strong>Fluent</strong> News · Spring 2006 23


HEALTHCARE<br />

CFD Assists Neonatal<br />

By Maciej K. Ginalski and Andrzej J. Nowak, Institute of Thermal Technology, Silesian University of Technology, Gliwice, Poland<br />

Pathlines, colored by temperature, illustrated the<br />

circulation inside the left side of the incubator<br />

Postprocessing courtesy of Michal / Nowak<br />

PREMATURE INFANTS generally enter the world with little protection from<br />

the harsh environment. Thermal comfort thus plays a crucial role in their survival<br />

and health. To provide the optimal environmental conditions for these infants,<br />

incubators are widely used. However, due to the complexity of the<br />

physical processes occurring within modern neonatal units during treatment<br />

procedures, comprehensive analysis involving the thermal comfort of premature<br />

infants can be difficult. The application of analytical calculations is currently not<br />

practical, and therefore numerical modeling is required.<br />

At the Silesian University of Technology in Gliwice, Poland, a CFD model of<br />

conjugate fluid flow and heat transfer in an infant incubator has been developed<br />

to support infant healthcare and improve medical equipment design. Accurate<br />

geometrical models representing the human body in a variety of postures are<br />

now easily created with high accuracy. Fluid flow analysis can be performed by<br />

taking into account mechanisms of heat transfer, such as radiation, convection,<br />

conduction, and even the evaporation of sweat and moisture from human skin.<br />

Since all infants, and particularly premature infants, are different from each other,<br />

the complex shape of the human body was generated using the CATIA package<br />

and described by a number of scaleable parameters. By changing the value of<br />

those parameters, a geometrical model can easily be adjusted to the individual<br />

24 <strong>Fluent</strong> News · Spring 2006


HEALTHCARE<br />

Intensive Care<br />

features of each analyzed patient or to his nursing position. The researchers used<br />

TGrid 4.0 to wrap the surface model of the infant’s body and to create the surface<br />

mesh to be exported to GAMBIT. In parallel, they created the geometry of<br />

the incubator using the CATIA package and combined the baby and incubator<br />

geometries using GAMBIT.<br />

Once a fine tetrahedral volume mesh was created, numerical calculations were performed<br />

in FLUENT. The research team enhanced the FLUENT model using several<br />

user-defined functions (UDFs) to control the processes of heat transfer together<br />

with internal heat generation inside the infant’s body, evaporation of moisture from<br />

the infant’s skin, and respiration. This last component of heat balance can be modeled<br />

as a steady-state or transient process. To make the calculation process fully<br />

automatic, managing software called MARCEL was developed to assist the team at<br />

all essential stages of the numerical simulation. MARCEL reads the provided data,<br />

manages the process of geometry and mesh creation, and sets the appropriate<br />

boundary and initial conditions. After receiving the results from the CFD solver,<br />

MARCEL then provides a basic report about the heat balance of the infant being<br />

analyzed. It is anticipated that a modified version of the program could be used by<br />

medical staff to perform heat balance calculations without possessing any knowledge<br />

of how to build and run CFD models.<br />

Contours of velocity magnitude on two planes through the<br />

incubator; red regions correspond to the flow inlets at the<br />

ends and outlets at the sides<br />

Validation of the code has been performed with comparison of the results to a<br />

series of clinical tests described in the medical literature. In all cases numerical<br />

models were created to represent analyzed infants and environmental conditions<br />

measured in the described tests. Results from the numerical simulations were<br />

compared with those obtained from the literature and proved to be very accurate.<br />

Researchers also performed a series of air temperature and velocity measurements<br />

inside an empty prototype of a new design. These measurements were<br />

fairly consistent with the results obtained from the numerical simulations. Based<br />

on this conclusion, modification of the ventilation system has been proposed and<br />

verified numerically.<br />

The incubator modeling project is still in the phase of testing and validation.<br />

However, even at the present stage, it is providing crucial information and is<br />

becoming an important tool in the treatment of premature babies at neonatal<br />

intensive care units. <br />

Suggested Reading:<br />

Ginalski, M., Nowak, A.J. and Wrobel, L.C.: Computational Model of Selected Transport<br />

Processes of the Premature Newborn Baby within an Infant Incubator. XXI International<br />

Congress of Theoretical and Applied Mechanics, Conference proceedings, Warsaw,<br />

Poland, August 2004.<br />

Ginalski, M., Nowak, A.J. and Wrobel, L.C.: Combined Heat and Fluid Flow in a Double<br />

Wall Infant Incubator. International Conference of Computational Methods in Sciences<br />

and Engineering, Conference proceedings, Loutraki, Greece, October 2005.<br />

Ginalski, M., Nowak, A.J. and Brandt, J.: Numerical Optimization of the Infant<br />

Incubator Ventilating System. Conference Proceedings, XIX National Congress of<br />

Thermodynamics, Sopot, Poland, September 2005.<br />

Acknowledgments<br />

The authors wish to express their gratitude to Nicolae Mera from the Center for<br />

Computational Fluid Dynamics at Leeds University, UK; Professor L.C. Wrobel from Brunel<br />

University, UK; Keith Hanna from <strong>Fluent</strong> Europe; and Michal / Nowak for their contribution<br />

and help in the project.<br />

Temperature contours show a uniform field in the region<br />

surrounding the infant<br />

Velocity vectors without (left) and with (right) an overhead plastic screen;<br />

the screen was found to help reduce radiative heat loss from the child<br />

<strong>Fluent</strong> News · Spring 2006 25


ENVIRONMENTAL<br />

Activated Sludge Basins<br />

Get on Track<br />

By Gregory Cartland Glover, Karim Essemiani and Jens Meinhold, Veolia Environment R&D Center, Maisons Laffitte, France,<br />

and Stephanie Vermande, Technical Division of Veolia Water, St. Maurice, France<br />

Hydrodynamic phenomena in<br />

the ps-ASB, showing gas fraction<br />

iso-surfaces (top) and pathlines<br />

of liquid velocity (bottom)<br />

concentration (mg/l)<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0<br />

Feed Aerobic Reaction<br />

experiment<br />

FLUENT ASM1 Anoxic Reaction<br />

WEST ASM1<br />

Ammonia<br />

Oxygen<br />

Nitrate<br />

30 60 90 120 150 180 210 240 270<br />

time (min)<br />

Profile plots of the reaction phenomena in the ps-ASB<br />

ACTIVATED SLUDGE BASINS (ASBs)<br />

are a key step in the treatment of municipal<br />

wastewater. ASBs are used to degrade<br />

biochemically reactive pollutants from<br />

water that is discharged into the natural<br />

environment. The bacteria found in the<br />

sludge consume and assimilate nutrients<br />

such as carbon, nitrogen, and phosphorous<br />

under several environmental conditions.<br />

The design of ASBs is difficult,<br />

however, due to factors including:<br />

• the treatment aims (pollutants<br />

to be degraded)<br />

• the influent wastewater composition<br />

(time and location<br />

dependent)<br />

• the application of liquid agitation<br />

and aeration (brushes,<br />

disks, membrane aerators,<br />

impellers, turbines and venturi<br />

tube mixers)<br />

• the location and spatial<br />

requirements (local weather,<br />

above or below ground,<br />

sparsely or densely populated<br />

regions)<br />

• the daily treatment loading<br />

(from a few hundred liters to<br />

thousands of cubic meters)<br />

Each one of these factors influences the<br />

size, form, and mode of operation (continuous<br />

or sequenced batch treatment) that<br />

is used when designing an ASB. Variation<br />

in some of the factors, such as daily<br />

changes in the temperature and in the<br />

influent composition and flow rate, can<br />

impact the efficacy of the treatment<br />

process. Thus, for the design of any ASB,<br />

an appropriate selection of agitators and<br />

aerators is essential to provide optimal<br />

environmental conditions for the treatment<br />

of wastewater.<br />

A study that was co-funded by Anjou<br />

Recherche, the R&D arm of Veolia Water,<br />

and the European Union was recently performed<br />

to develop a computational strategy<br />

for modeling ASBs. The primary focus<br />

of the study was to examine the effect of<br />

the local hydrodynamics on the biochemical<br />

reactions observed in ASBs for both<br />

pilot and real scale processes. The results<br />

provided an improved understanding of<br />

the hydrodynamic impact on the environmental<br />

conditions experienced by the<br />

bacteria.<br />

A pilot-scale ASB (ps-ASB) was used to<br />

establish a database of hydrodynamic data<br />

(circulation velocity and mean bubble<br />

sizes) and interphase mass transfer data<br />

under different operating conditions. The<br />

ps-ASB was then used to study the impact<br />

of the normal and abnormal operating<br />

conditions on the biochemical reaction<br />

phenomena by monitoring the change<br />

in concentration of key wastewater<br />

constituents, such as ammonia, organic<br />

carbon, nitrates, organic nitrogen, and<br />

particulate matter.<br />

The ps-ASB applied different environmental<br />

conditions in a programmed sequence.<br />

The different conditions were affected by<br />

contacting the mixed liquor, a mixture of<br />

wastewater and activated sludge, with air<br />

in the form of fine bubbles, and applying<br />

a circulation velocity with a marine-type<br />

impeller. This condition enabled the aerobic<br />

oxidation of both organic carbon and<br />

nitrogen (in the form of ammonia) to carbon<br />

dioxide and nitrate, respectively. A<br />

second condition only agitated the biomass<br />

in the mixed liquor to reduce the<br />

overall nitrogen content of the wastewater.<br />

Under these conditions a different<br />

type of biomass (anoxic bacteria) consumes<br />

the nitrate producing nitrogen,<br />

which then evolves as a gas.<br />

The operational data was used to create a<br />

26 <strong>Fluent</strong> News · Spring 2006


ENVIRONMENTAL<br />

hydrodynamic and biochemical reaction model of<br />

the ps-ASB. The first step was to model the biochemical<br />

reactions assuming perfectly mixed reactor<br />

conditions with the wastewater modeling tool<br />

WEST ® [1], which uses the Activated Sludge Model<br />

No. 1 (ASM1) protocol [2].<br />

The mixed liquor composition and the calibrated<br />

kinetic and stoichiometric parameters derived from<br />

this global model were then applied to a converged<br />

local hydrodynamic solution constructed in FLUENT.<br />

The Eulerian multiphase model was employed with<br />

the dispersed phase form of the k-ε turbulence<br />

model to resolve the gas-liquid motion under steady<br />

flow conditions.<br />

The application of the ASM1 protocol to FLUENT<br />

was validated by analysis of the component profiles<br />

during the aerobic and anoxic reaction phases. The<br />

rates under the different operating conditions were<br />

accurate to within 10% of the experimentally measured<br />

reaction rates, where the accepted error was<br />

10 to 15%.<br />

Liquid velocity pathlines and volume fraction contours<br />

of the ps-ASB were used to illustrate the<br />

hydrodynamics of the basin. The distribution of dissolved<br />

oxygen after one hour of aeration was also<br />

examined. The difference between the cold and<br />

warm iso-surfaces was found to be 0.4 mg/l over<br />

the range from 2.9 to 3.4 mg/l. At this point in the<br />

simulation, the oxygen concentration gradient does<br />

not influence which biochemical reaction processes<br />

are predominant. However, at a mean concentration<br />

of less than 0.5 mg/l different treatment<br />

regimes could be found in the reactor (informal<br />

aerobic and anoxic zones).<br />

The aeration and agitation regimes applied to the<br />

ps-ASB were considered to be perfectly mixed and at<br />

a comparatively small scale for treatment processes.<br />

Thus, the effects observed with the ps-ASB may be<br />

different from those of real scale Activated Sludge<br />

Basins (rs-ASBs) where volumes of up to 10,000m 3<br />

are widely used. At the larger scale, perfectly mixed<br />

reactor or plug flow regimes are not guaranteed,<br />

and the potential influence of the mixing regimes is<br />

far greater and more difficult to categorize. Thus,<br />

incorrect operating conditions will cause mixing<br />

regimes that would hinder the ability of the ASB to<br />

meet the treatment objectives.<br />

To test this possibility, two rs-ASBs with different<br />

geometries and hydrodynamic regimes were simulated.<br />

The basin rs1-ASB is a carrousel type ASB, whereas<br />

rs2-ASB is a race-track type ASB. The information<br />

derived from these simulations (circulation velocity<br />

and oxygen mass transfer characteristics) indicated<br />

how well each ASB could meet its objectives.<br />

The pathlines for rs1-ASB indicate that the conditions<br />

are near optimal, while pathlines in rs2-ASB<br />

showed that the process was not close to optimal<br />

conditions. The flow field in rs2-ASB was characterized<br />

by low horizontal liquid velocities and large<br />

recirculation zones and short-circuiting upstream of<br />

the bubble plumes. The liquid circulation patterns<br />

were disrupted by the bubble plumes, causing them<br />

to degrade. In general, poor liquid circulation<br />

reduces the oxygen mass transfer rates and the dispersion<br />

of oxygen throughout the ASB. The ability<br />

of the biomass in this particular basin to aerobically<br />

treat the wastewater could therefore be limited by<br />

the operating conditions applied.<br />

Further simulations were performed to improve the<br />

aeration strategy by analyzing the effect that different<br />

aerator and agitator configurations had on these<br />

two design parameters. The simulations examined<br />

the positions of the aerators and agitators as well as<br />

the agitation rate and size of the impellers, with the<br />

single goal of improving the mass transfer characteristics<br />

of rs2-ASB. The hydrodynamics of the modified<br />

design showed improved liquid circulation<br />

with less disruption of the flow by the bubble<br />

plumes, resulting in a 250% increase of the circulation<br />

velocity.<br />

The computational strategies developed during the<br />

course of this study have enabled Veolia Water<br />

to obtain a numerical tool that can aide decisions<br />

made in selecting aeration and agitation configurations<br />

for Activated Sludge Basins, for the design of new<br />

technologies and retrofitting of existing installations. <br />

References:<br />

1 Hemmis, N.V., WEST ® , Hemmis N.V., Kortrijk, Belgium,<br />

2003.<br />

2 Henze, M., Gujer, W., Mino, T., and van Loosdrecht, M.:<br />

Scientific and Technical Report No. 9. IWA Publishing,<br />

London, UK, 2000.<br />

rs1-ASB rs2-ASB Modified rs2-ASB<br />

Hydrodynamic phenomena in the rs-ASB, showing gas fraction iso-surfaces (top) and pathlines of liquid velocity (bottom)<br />

<strong>Fluent</strong> News · Spring 2006 27


ACADEMIC NEWS<br />

A Dry Passage<br />

to the Afterlife<br />

By Essam E. Khalil and Omar A.A. AbdelAziz,<br />

Cairo University, Egypt<br />

EGYPT WAS ONCE HOME to the Pharaohs,<br />

one of the oldest and most sophisticated civilizations<br />

in the ancient world. Many Egyptian artifacts,<br />

treasures, and buildings are a priceless part of our<br />

world heritage. The Valley of the Kings at Luxor is<br />

a unique and world famous site where thousands<br />

of years ago, many Egyptian monarchs had elaborate<br />

tombs built to ensure their safe passage into<br />

the afterlife. Many of these tombs were robbed<br />

after being discovered, but others have survived<br />

and their greatest remaining treasures are the<br />

beautiful, vivid, yet fragile wall paintings and decorations<br />

that lie within. Each year, tourists flock to<br />

Egypt to see these awe-inspiring tombs, but their<br />

visits cause a major problem for curators. The heat<br />

and humidity given off by visitors in the enclosed<br />

chambers and passageways cause damage to the<br />

plasterwork and paintings. Hence, the ventilation<br />

system and resultant air flow patterns in a tomb<br />

are critical to the preservation of the exhibits so<br />

that they may be kept open to tourists for many<br />

years to come.<br />

The Egyptian government and the Supreme<br />

Council of Antiquities approached ventilation<br />

experts at Cairo University for help in solving this<br />

problem for the archeaological tombs in the Valley<br />

of the Kings. The engineers were asked to devise a<br />

climate control system for these tombs; a pilot<br />

study was conducted on the tomb of Ramses VII.<br />

Using GAMBIT, a model of approximately one<br />

million cells was created that yielded good geometric<br />

representation of the tomb passageway<br />

with the sarcophagus in the main room and a<br />

large number of visitors. They then used custom<br />

models in FLUENT to simulate human breathing<br />

and heat generation, and considered a worst case<br />

scenario when 20 adult visitors are in the tomb at<br />

the same time. Parametric CFD simulations were<br />

performed to predict what the relative humidity, a<br />

key variable, would be like near the wall paintings<br />

for a given air extraction system design and tomb<br />

vent locations. Their modeling work was used to<br />

determine the optimized air flow pattern for the<br />

tomb. Based on their findings, the installation of<br />

raised flow exhausts that were unobtrusive to<br />

visitors was recommended. The tomb now has<br />

minimal adverse flow and humidity gradients<br />

within the chamber, thereby helping to preserve<br />

this priceless piece of history for future generations<br />

of visitors. <br />

Predicted local humidity levels on a plane near the<br />

main wall paintings of the tomb of Ramses VII with a<br />

representative adult male tourist figure; the<br />

sarcophogus is in the foreground<br />

Postprocessing courtesy of Maciej Ginalski<br />

Convective<br />

By Daniele Melideo, Davide Mazzini, Enrico<br />

A NUMBER OF BENEFITS have resulted from<br />

recent reductions in weight and improvements to<br />

the efficiency of aeronautical gearboxes. Reduced<br />

heat generation inside the gearbox means that less<br />

oil is needed to maintain a certain transmission<br />

operating temperature. In addition, reductions in<br />

surge system losses allow the lubrication system<br />

components, such as tanks, high pressure tubes,<br />

filters, and pumps to be smaller in size.<br />

28 <strong>Fluent</strong> News · Spring 2006<br />

Predicted pathlines colored by moisture<br />

in the air in the tomb of Ramses VII;<br />

the flow is from left to right<br />

Postprocessing courtesy of Maciej Ginalski<br />

In the framework of thesis work carried out in the<br />

Department of Mechanics, Nuclear and Production<br />

Engineering (DIMNP) of the University of Pisa, the<br />

convective motion of the air inside a gearbox<br />

induced by the rotation of the gear drive components<br />

has been investigated. FLUENT was used for<br />

the numerical modeling component of the project,<br />

which focused on a single gear pair. The reference<br />

equipment is located at the Research Center on


ACADEMIC NEWS<br />

Student Submariners Peddle<br />

Their Way to Victory<br />

By Keith Hanna, <strong>Fluent</strong> News and Joost Sterenborg, Technical University Delft, Netherlands<br />

MANY UNIQUE CFD applications<br />

come across the desk of <strong>Fluent</strong><br />

News, and one that recently appeared<br />

was of a human powered submarine.<br />

Students from university engineering<br />

teams around the world have come<br />

together for an international humanpowered<br />

submarine competition for<br />

the last eight years. The teams of student<br />

engineers have to design their<br />

own submarines from the ground up<br />

using modern computer-based tools,<br />

and manufacture and test their<br />

designs for performance and safety.<br />

They also need to secure enough<br />

sponsorship funding to cover the cost<br />

of their team, including travel to the<br />

competition venue.<br />

WASUB, an intrepid team from the<br />

Technical University of Delft in the<br />

Netherlands approached <strong>Fluent</strong><br />

Benelux, along with other companies<br />

in the region, for sponsorship last<br />

year. They used Pro/ENGINEER ®<br />

from PTC to design their submarine<br />

in CAD, and FLUENT to hone the<br />

hull and fin shapes for optimal hydrodynamic<br />

drag characteristics. After<br />

extensive testing of the WASUB<br />

in the MARIN towing tank in<br />

Wageningen, the craft was shipped<br />

to America for the races. The team<br />

came away with several prestigious<br />

prizes, including best design and<br />

fastest in two categories. Throughout<br />

the process, the young engineers<br />

gained modern product design engineering<br />

experience in a competitive<br />

situation. <strong>Fluent</strong> salutes these brave<br />

Dutch submariners and their revolutionary<br />

WASUB design! <br />

More.info@<br />

http://wasub.oli.tudelft.nl<br />

The WASUB Team in Maryland, June 2005<br />

CAD layout of WASUB for a single<br />

human-powered propulsion system<br />

Axisymmetric FLUENT prediction of static<br />

pressure around the WASUB hull shape<br />

Motions Inside a Gearbox<br />

Manfredi, and Maria Vittoria Salvetti, The University of Pisa, Italy<br />

Advanced Technology Mechanical Transmissions<br />

(CRTM) of DIMNP, as part of a collaboration with<br />

the Italian aerospace company Avio. The test section<br />

consists of two rotating shafts that support the test<br />

gears. Wheel lubrication is ensured by two spray<br />

bars injecting oil upstream and downstream of the<br />

mesh point, where the gear wheels come together.<br />

The first activity conducted was to test different<br />

turbulence models (Spalart-Allmaras, k-ε, k-ω, and<br />

Reynolds stress). Their performance was evaluated<br />

by studying 2D cases of Couette motion, consisting<br />

of two coaxial cylinders rotating at different velocities.<br />

The k-ω model was chosen as the most costeffective<br />

turbulence model, since it was able to<br />

capture the swirling velocity component better than<br />

the other RANS models, and did not require the<br />

additional computational resources of RSM. This<br />

model was used for subsequent numerical studies.<br />

The 3D simulation of the gearbox focused on the air<br />

motion induced by the wheel rotation, and included<br />

the effects of protruding hardware, such as bolt<br />

heads. Owing to a plane of symmetry, half of the<br />

gearbox was modeled. The oil presence was neglected<br />

in the first round of simulations. The wheel<br />

rotation was simulated using the multiple reference<br />

frames (MRF) model. The gear teeth, small on the<br />

scale of the wheels, were neglected. The bolt heads<br />

were found to entrain the air creating a venting<br />

effect. While the wheel rotation produced mostly<br />

circumferential flow, a significant axial velocity was<br />

observed between the spray bars and the mesh<br />

point, suggesting a possible deflection of the oil jet.<br />

In the next stage of the work, a more refined grid<br />

and the use of the sliding mesh model will be<br />

adopted, and the oil presence will be taken into<br />

account. <br />

Axial velocity on the symmetry<br />

plane, with vectors showing the<br />

axial flow near the mesh point<br />

<strong>Fluent</strong> News · Spring 2006 29


PRODUCT NEWS<br />

<strong>Fluent</strong> & Microsoft Team to Deliver<br />

64-bit FLUENT on Windows Clusters<br />

By Diana Collier, Barbara Hutchings, and Rongguang Jia, <strong>Fluent</strong> Inc.<br />

ing the way for cluster-based FLUENT simulations for Windows<br />

users. With this 64-bit support, Windows users will now be able to run<br />

much larger simulations than currently feasible and they will also see<br />

performance improvements due to the enhanced memory management<br />

and wider memory bandwidth available with 64-bit processors.<br />

This good news for Windows users is the result of close collaboration<br />

between <strong>Fluent</strong> and Microsoft. With excellent technical support from<br />

Microsoft, <strong>Fluent</strong>’s development team has optimized FLUENT 6.3 on<br />

CCS. The resulting cluster solution uses the Microsoft MPI (Message<br />

Passing Interface) software layer for data communication between<br />

processors on the cluster, and supports a variety of interconnect<br />

options including Gigabit Ethernet (GigE), Infiniband, and Myrinet.<br />

FLUENT 6.3 also takes advantage of the Microsoft job scheduler that<br />

ships with CCS, providing an off-the-shelf solution for launching and<br />

controlling jobs on the cluster.<br />

Water ski jumper simulation showcased on the 64-bit Windows cluster at SC05<br />

Courtesy of Sports Engineering Group, Sheffield Hallam University; postprocessed using EnSight from CEI<br />

speedup<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

0<br />

FL5L1<br />

FL5L2<br />

FL5L3<br />

Linear<br />

2 4 6 8 10<br />

number of processors<br />

Parallel performance of 64-bit FLUENT running the<br />

FL5L1, FL5L2, and FL5L3 large benchmark cases,<br />

using Windows CCS 2003 on 3.4 GHz Intel EM64T<br />

dual-CPU nodes connected with gigE<br />

30%<br />

28.5%<br />

25%<br />

22.0%<br />

20%<br />

16.7%<br />

14.3%<br />

15%<br />

11.8%<br />

12.5%<br />

15.3%<br />

15.7%<br />

10%<br />

5%<br />

0%<br />

15s1 15s2 fl5s3 fl5m1 15m2 15m3 fl5l1 fl5l2<br />

performance boost<br />

Serial performance boost with the 64-bit solution,<br />

relative to 32-bit, for the small (FL5S1 and FL5S2),<br />

medium (FL5M1, FL5M2, and FL5M3), and large<br />

(FL5L1 and FL5L2) benchmark cases<br />

THE AVAILABILITY OF 64-BIT<br />

off-the-shelf computing from<br />

Intel and AMD has provided<br />

great value to <strong>Fluent</strong> customers<br />

seeking to run larger, more<br />

memory intensive CFD simulations.<br />

With the upcoming<br />

release of FLUENT 6.3, 64-bit<br />

will become an option to customers<br />

running the Microsoft ®<br />

Windows ® operating system.<br />

Many FLUENT customers are<br />

already running a 32-bit version<br />

of FLUENT on the 64-bit<br />

Windows XP operating system,<br />

but in this configuration FLUENT<br />

does not take advantage of<br />

extended memory addressing.<br />

With FLUENT 6.3, full 64-bit<br />

capability will be supported on<br />

both desktop and server systems<br />

running Windows. On the server<br />

side, FLUENT 6.3 has been ported<br />

and optimized to run under<br />

the new Microsoft ® Windows ®<br />

Compute Cluster Server 2003<br />

(CCS) operating system, open-<br />

Benchmarking<br />

The performance of 64-bit FLUENT under CCS is excellent. Speed<br />

improvements for serial performance, relative to the 32-bit FLUENT<br />

version, are in the range of 10-30% for many of the examples in<br />

<strong>Fluent</strong>’s standard benchmark problem set. Parallel scaling is also quite<br />

good. Initial scaling studies using the larger benchmarks on a Windows<br />

CCS cluster are on par with what is observed on similarly configured<br />

clusters running Linux. For example, on an 8-CPU cluster connected<br />

with GigE, a speed-up of 6.4 was obtained using the FL5L2 benchmark.<br />

Supercomputing 2005<br />

Microsoft showcased their new high-performance computing platform<br />

at Supercomputing 2005 (SC05), held in Seattle November 14-18,<br />

2005. FLUENT was proud to be part of the Microsoft exhibit, demonstrating<br />

FLUENT 6.3 running as a 64-bit application under Windows<br />

Compute Cluster Server 2003 on a 16 CPU cluster provided by<br />

IBM and using a Myrinet interconnect. The showcased simulation was<br />

a multi-million cell model of a water ski jumper, courtesy of Sports<br />

Engineering Group, Sheffield Hallam University. In fact, several <strong>Fluent</strong><br />

partners at SC05 demonstrated FLUENT simulations running on CCS<br />

clusters, including Dell, Mellanox, Voltaire, and Broadcom. The widespread<br />

success of these demonstrations, using pre-release versions of<br />

both the operating system and FLUENT, was a confirmation of<br />

Microsoft’s stated goal to provide HPC solutions that are easy to<br />

deploy, operate, and integrate with existing infrastructure and tools. <br />

For information on obtaining FLUENT 6.3 beta for Windows Compute Cluster<br />

Server 2003, contact your local <strong>Fluent</strong> account manager.<br />

More.info@<br />

www.microsoft.com/windowsserver2003/<br />

ccs/overview.mspx<br />

30 <strong>Fluent</strong> News · Spring 2006


PRODUCT NEWS<br />

Quick Turnaround with<br />

Rapid Flow Modeling<br />

By André Bakker, FloWizard Product Manager, and Laurent Collonge, FLUENT for CATIA V5 Product Manager<br />

THE OBJECTIVE OF RAPID FLOW MODELING is<br />

to bring CFD to the broader engineering design<br />

market, by providing fluid flow analysis tools that let<br />

engineers quickly turn their CAD models into CFD<br />

results. There are now two full-fledged rapid flow<br />

modeling products available in <strong>Fluent</strong>’s product<br />

line: FloWizard and FLUENT for CATIA V5. With<br />

these products, <strong>Fluent</strong> is taking the lead in what is<br />

expected to be a fast growing segment of the CFD<br />

software market.<br />

Many recent technology advances have made rapid<br />

flow modeling possible. The increased speed of<br />

computer hardware has reduced the calculation<br />

time for many CFD applications to fit within engineering<br />

design timeframes. A simulation that previously<br />

might have taken tens of hours on a workstation<br />

can now be run in well under an hour on a<br />

standard PC. On the software side, it is now possible<br />

to embed the CFD experience in highly automated<br />

products that significantly reduce the turnaround<br />

time for a fluid flow analysis.<br />

Modern companies have an increasing need for engineering<br />

analysis to be tightly integrated into their<br />

design and PLM (product lifecycle management)<br />

processes. This requires software that interfaces well<br />

with CAD and other engineering software tools. It<br />

also means that the fluids analysis software should be<br />

accessible to engineering designers, who may have<br />

only intermittent needs to run simulations. FloWizard<br />

and FLUENT for CATIA V5 share characteristics that<br />

allow them to meet these needs. They each have a<br />

high level of connectivity with CAD or PLM products.<br />

This is important, since most of their users are engineers<br />

involved in product design. FloWizard accepts<br />

a wide variety of CAD and mesh file formats, and<br />

has tight CAD connections with SolidWorks ® ,<br />

Pro/ENGINEER ® and UGS’ NX. FLUENT for CATIA V5<br />

embeds the same rapid flow modeling technology<br />

deep into the CATIA V5 PLM system.<br />

Both packages focus on modeling well understood<br />

physics, including compressible and incompressible<br />

fluid flow and heat transfer, both laminar and turbulent.<br />

Stationary equipment can be modeled as well<br />

as rotating machinery. Tasks that often require a lot<br />

of user interaction, such as geometry cleanup,<br />

meshing, and solving are fully automated. Not only<br />

does this save time, it also leads to a significantly<br />

reduced learning curve.<br />

Both products make use of the fast, accurate, and<br />

well-validated FLUENT 6 solver in the background.<br />

As an additional benefit, the files generated by<br />

FloWizard and FLUENT for CATIA V5 can be shared<br />

with FLUENT users as needed. This also helps<br />

improve the level of collaboration between the<br />

analysis and design teams. Furthermore, FloWizard<br />

Starting with a pump model in SolidWorks, engineers can quickly send their<br />

model to FloWizard. There it will be meshed and solved using FloWizard’s<br />

automated tools. The final results can be used by the engineer to decide<br />

upon possible design changes, thus closing the full design cycle.<br />

offers built-in collaboration tools that let multiple<br />

users simultaneously connect to the same session<br />

for design reviews. When there is a need to model<br />

large problems, both products offer the ability<br />

to run calculations in parallel on a local or remote<br />

network.<br />

In many companies, there is also a need to efficiently<br />

model recurring problems related to their product<br />

lines. For this purpose, custom tools can be built<br />

upon the rapid flow modeling platform. FloWizard is<br />

fully customizable using the Python programming<br />

language which has been used by several companies<br />

to build custom, organization specific analysis<br />

tools. <br />

<strong>Fluent</strong> News · Spring 2006 31


PRODUCT NEWS<br />

High quality Cooper and tet meshes in a redesigned adaptorconnector<br />

originating from a Pro/E part file<br />

Impressive Prepro<br />

By Erling Eklund, GAMBIT and TGrid Product Manager<br />

THE SPRING RELEASES of several preprocessing products from <strong>Fluent</strong><br />

promise many exciting new features and capabilities, ranging from CAD<br />

integration to wrapping technology.<br />

GAMBIT<br />

In GAMBIT 2.3 the focus continues to be on CAD import, geometry manipulation,<br />

and advanced meshing. A new CATIA V5 translator has been added, as well<br />

as new CAD Connections for SolidWorks ® , Pro/ENGINEER ® , and UGS’ NX.<br />

These Connections use native surface, connectivity, and topology information,<br />

and eliminate the need to use a third party generic CAD format to translate your<br />

geometry model. With the CAD Connections, tools developed by <strong>Fluent</strong> reside<br />

in the CAD packages to perform a number of integrity checks prior to export,<br />

which guarantees that clean CAD integration is achieved.<br />

All edge and face construction tools in GAMBIT 2.3 can be utilized on these<br />

CAD geometries including the new closest point projection. In addition,<br />

design changes and general mesh-to-CAD conversion can be made using a<br />

new Advanced Covering algorithm. The cleanup tools continue to expand<br />

with a new sliver face removal capability.<br />

A surface mesh of occupants<br />

in a truck cabin, created<br />

using the new surface<br />

wrapper in TGrid<br />

In the area of meshing, the size function capability has been improved<br />

through speed-ups and added functionality, such as individual minimum<br />

size assignments, 2D proximity, size function blending, and a new, exciting<br />

meshed size function, starting on a boundary layer cap.<br />

For boundary layers, a new first height, last-ratio-based cell growth function<br />

has been added, and the full 3D boundary layer can be examined prior to<br />

volume meshing. A new quad split option in tetrahedral meshing allows for<br />

combined hex and tet meshing, without the need for pyramids, which in<br />

many cases leads to significant cell quality improvements.<br />

TGrid<br />

The most important new feature in TGrid 4.0 is the surface wrapper, which<br />

allows connectivity, defeaturing, and surface meshing to be handled in a<br />

single operation. Once the large holes have been covered, the tool wraps<br />

the geometry with a high-quality surface mesh from which a volume mesh<br />

can then be created. Gaps, slivers, and overlaps in the geometry are automatically<br />

cleaned-up in the process. The advanced wrapping procedure<br />

provides a full spectrum of size controls, including local and global<br />

32 <strong>Fluent</strong> News · Spring 2006


PARTNERSHIPS<br />

Discrete Element<br />

Modeling of<br />

Particles for FLUENT<br />

By Peter Weitzman and John Favier, DEM Solutions, Edinburgh, UK<br />

The 3Matic-for-<strong>Fluent</strong> interface, showing all three levels of operations:<br />

Manual (buttons on top), Autofix (right panel), and Batch Fix (left panel)<br />

cessing<br />

proximity, curvature, and zone-based size functions. It also supports<br />

controlled size growth, and unique coarsening and feature capturing capabilities.<br />

Other functionalities include the ability to fully examine the region<br />

to be wrapped, prior to the wrapping procedure, with a guaranteed hole<br />

finder. There are local and global feature extraction capabilities, automatic<br />

fixing of incorrect topology, and automatic mesh quality improvement.<br />

Using a strategy based on an “over-refined” initial mesh, imprinting and<br />

coarsening together produce the highest-quality surface mesh on the<br />

market today that is based on wrapper technology.<br />

In addition to the introduction of the surface wrapper, the prism layer<br />

technology in TGrid 4.0 has been expanded. The initial normal growth direction<br />

has been improved, leading to overall speed and quality improvements.<br />

Two automated functions have been added: one to detect prism layer proximity/collision,<br />

which includes automatic prism height adjustment, and another<br />

to detect and peal off prism layers in sharp corners, which includes automatic<br />

non-conformal re-meshing of prism sides. The ability to handle prism<br />

layer growth on double-sided, zero-thickness walls has also been added.<br />

DEM SOLUTIONS is the developer of EDEM, discrete element method<br />

(DEM) software for particle flow simulation. EDEM is used to simulate the dynamics<br />

of solid particles including inter-particle and particle-wall collisions.<br />

<strong>Fluent</strong> and DEM Solutions have jointly developed a new plug-in for FLUENT that<br />

can be used to simulate the dynamics of multiphase fluid-particle systems. The<br />

EDEM plug-in complements the multiphase models already available in FLUENT<br />

by adding the ability to include particle attributes such as shape and size distribution,<br />

material properties, and initial placement, and to obtain detailed information<br />

on collisions, clumping, adhesion to boundary surfaces, and exchange<br />

with the surrounding fluid. In addition to the fluid drag, other forces such as<br />

gravity, electrostatic, electromagnetic, bonding, and cohesion can be applied to<br />

the particles. An open interface allows additional user-defined forces and<br />

coupling with other physical models.<br />

The EDEM plug-in can provide important information on multiphase flows across<br />

a broad range of industries and applications. It is useful in applications where the<br />

particles must be transported efficiently, without damage or excessive buildup,<br />

or when particle-equipment interactions are critical. Some examples are the<br />

optimization of pneumatic transport systems, fluidization, filtration, prevention<br />

of sand damage, and reactor transport.<br />

The EDEM software can be used by itself to model single-phase systems of solid<br />

particles. Some examples of single phase applications include bulk material handling<br />

in conveyors, chutes and hoppers, rock drilling, excavation, agricultural<br />

machinery, and various pharmaceutical manufacturing processes. <br />

A large number of enhancements have also been added in the areas of<br />

surface mesh intersection, re-meshing, local manipulation, and more.<br />

3Matic-for-<strong>Fluent</strong><br />

The wrapping of highly complex industrial models often involves dealing<br />

with hundreds of files, all of different quality and origin. In collaboration<br />

with Materialise, a new product has been developed to meet this type of<br />

challenge. 3Matic-for-<strong>Fluent</strong> 1.0 is designed to handle all types of CAD<br />

formats (native, standard, and faceted) and produce a coherent, optimal,<br />

and clean faceted mesh that is ready to wrap in TGrid 4.0. This customized<br />

tool allows data to be handled in three ways. First, connectivity, cleanup,<br />

hole filling, and faceting of a single CAD file can be investigated, step-bystep,<br />

so that the impact of parameter settings can be better understood.<br />

Second, some of the files can be run through an auto-fix sequence to evaluate<br />

consistent behavior through different CAD formats. Third, hundreds<br />

of mixed CAD files can be run through the batch converter using the<br />

selected parameters, all producing TGrid mesh files. <br />

3Matic-for-<strong>Fluent</strong> is sold and distributed by <strong>Fluent</strong>. Please contact your local <strong>Fluent</strong><br />

office for more information.<br />

Entrainment of particles in an air-stream<br />

using FLUENT-EDEM co-simulation<br />

More.info@<br />

www.dem-solutions.com<br />

<strong>Fluent</strong> News · Spring 2006 33


SUPPORT CORNER<br />

Mapping Thermal Data<br />

from FLUENT to Structural<br />

Codes Quickly<br />

By Aleksandra Egelja-Maruszewski, <strong>Fluent</strong> Inc.<br />

“<strong>Fluent</strong>’s add-on UDF for temperature<br />

mapping is an excellent tool. It allows<br />

quick and accurate interpolation of<br />

temperatures between a CFD model<br />

and a structural mesh. It has cut the<br />

time and effort it takes us to map<br />

temperature data from FLUENT to<br />

FEA drastically. It is simple to use<br />

and flexible. If you need to map<br />

temperatures from a CFD model<br />

to a structural mesh, this is the tool<br />

to use.”<br />

– William Towne<br />

CAE Engineer<br />

Performance Prediction<br />

Mercury Marine<br />

IN MANY MODERN PROCESSES TODAY there is<br />

an increased need to solve multiphysics problems,<br />

such as thermal coupling. Some of these problems<br />

can be solved with one simulation software only;<br />

others require the combination of multiple software<br />

packages. In extreme thermal environments, there<br />

are various types of thermal loads on components<br />

and structures. Applying these thermal loads correctly<br />

is one of the first steps required for achieving<br />

a reliable thermal stress analysis of the component.<br />

Thermal loads can be applied as a temperature<br />

distribution within the solid, or as heat transfer<br />

coefficients on the skin of the solid where the heat<br />

transfer takes place.<br />

Most CFD tools have advanced fluid dynamics and<br />

heat transfer algorithms, but do not have built-in<br />

advanced solid mechanics analysis capabilities such<br />

as thermal stress analysis and fatigue prediction. On<br />

the other hand, most of the advanced finite element<br />

analysis (FEA) tools have advanced solid mechanics<br />

algorithms, but do not offer the necessary advanced<br />

fluid dynamics and heat transfer capabilities to<br />

determine the correct thermal loads for the thermal<br />

stress analysis. For scenarios involving fluid flow and<br />

thermal loading, the two simulation tools become<br />

dependent on each other.<br />

To make it easier to perform combined simulations<br />

such as these, engineers at <strong>Fluent</strong> have developed a<br />

user-defined function (UDF) for coupling CFD and<br />

FEA for thermal stress analysis. With this new capability,<br />

FLUENT is used to solve for the flow and thermal<br />

analysis in the fluid and solid zones of a part or<br />

component. The temperature data of the solid<br />

region is then supplied to the FE solver as the thermal<br />

load profile for the thermal stress analysis.<br />

Because the CFD mesh is typically much finer than<br />

the FE mesh, it is necessary to interpolate, or map,<br />

the CFD results to the FE mesh. The mapping is<br />

accomplished through a Scheme file (read into<br />

FLUENT) and the UDF.<br />

For the case of exporting data from FLUENT to<br />

ABAQUS, for example, the coupled solution would<br />

proceed as follows.<br />

1. The user solves the case in FLUENT.<br />

2. The user reads the Scheme file and<br />

compiled UDF into FLUENT.<br />

3. The user reads the ABAQUS ASCII format<br />

mesh into FLUENT. The FLUENT UDF<br />

uses import filters to read the nodes,<br />

elements, and connectivity from the<br />

ABAQUS mesh.<br />

34 <strong>Fluent</strong> News · Spring 2006


SUPPORT CORNER<br />

The panel used for input to the mesh interpolator<br />

4. For each FE node/element, the UDF finds<br />

the closest solution location in the CFD<br />

mesh. A binary space partitioning (BSP)<br />

tree is used to find the closest node<br />

by recursively subdividing the space<br />

containing the geometry, making fast<br />

geometric searching possible.<br />

5. The UDF interpolates from the closest<br />

node.<br />

6. The UDF instructs FLUENT to export the<br />

results in a format suitable for ABAQUS.<br />

There are two methods for performing this<br />

mapping procedure:<br />

• The temperature data inside the solid<br />

region is exported. For each node in the<br />

FE mesh of the solid region, temperature<br />

data will be written to an output file.<br />

• The temperature and heat transfer coefficient<br />

at the walls that are solid/fluid<br />

interfaces are exported. For each surface<br />

element of the FE mesh, the temperature<br />

and heat transfer coefficient will be<br />

written to an output file.<br />

The UDF is not designed to manage strongly coupled<br />

CFD and FEA calculations that require many<br />

iterations back and forth between the solvers.<br />

Instead, it is designed to make individual data transfers<br />

easier to do. The UDF, which has panels for<br />

input, is currently available for ABAQUS, ANSYS,<br />

and PATRAN formats, but it can easily be<br />

customized for other structural analysis codes. The<br />

mesh can be built in millimeters, centimeters, inches,<br />

or meters. Temperatures can be in Kelvin,<br />

Fahrenheit, or Celsius. The UDF performs in serial<br />

and parallel versions of FLUENT. Data for different<br />

solid zones or wall zones can be exported separately<br />

by reading the individual mesh files of the zones.<br />

With additional customization, other properties can<br />

be exported as well.<br />

The compiled libraries are freely available for<br />

exporting steady-state data. Simply download the<br />

CFD/FEA Thermal Coupling UDF at no charge from<br />

the UDF examples section of the UDF archive on the<br />

User Services Center at www.fluentusers.com. (You<br />

will need your username and password to log on<br />

to the site.)<br />

To customize the UDF or obtain libraries for the<br />

export of transient or time-averaged data, or if you<br />

have questions regarding the use of this UDF, please<br />

contact your local <strong>Fluent</strong> office or log a help<br />

request through the Online Technical Support<br />

Portal, accessible from the User Services Center:<br />

http://clarify.fluent.com/eSupport<br />

The geometry and mesh of the powerhead of a<br />

marine outboard is used to illustrate this capability.<br />

Using the UDF, a 2.8 million cell FLUENT mesh<br />

was mapped to a 430,000 element ABAQUS mesh<br />

in 14 seconds. Contour plots of static temperature<br />

computed by FLUENT (left) and ABAQUS (right) are<br />

shown for several views of the powerhead assembly<br />

Courtesy of Mercury Marine<br />

<strong>Fluent</strong> News · Spring 2006 35


AROUND FLUENT<br />

<strong>Fluent</strong> Opens Larger<br />

Office in Ann Arbor, MI<br />

FLUENT HAS OPENED a new, larger office in<br />

Ann Arbor, MI. This expanded office location allows<br />

for the addition of more customer service staff,<br />

an expanded training facility, and a much larger<br />

computer resource center. The new training facility<br />

can accommodate more people and allows for the<br />

addition of new educational sessions on advanced<br />

automotive application areas such as underhood<br />

thermal management, noise prediction, and HVAC<br />

system analysis.<br />

Upcoming User<br />

Group Meetings<br />

36 <strong>Fluent</strong> News · Spring 2006<br />

<strong>Fluent</strong> invites you to take full advantage of the<br />

facility. Meet with the US customer services team,<br />

get trained on new product innovations, and<br />

discuss how the Ann Arbor team can provide consulting<br />

services to meet your engineering needs.<br />

<strong>Fluent</strong><br />

3005 Boardwalk, Suite 100<br />

Ann Arbor, MI 48108<br />

734 213 6821<br />

US CFD Summit May 22 – 24<br />

US Automotive CFD Summit Sept. 11 – 12<br />

<strong>Fluent</strong> Sweden Sept. 18 – 19<br />

<strong>Fluent</strong> Europe Sept. 20 – 22<br />

<strong>Fluent</strong> Benelux Oct. 4 – 6<br />

<strong>Fluent</strong> China Nov. 6 – 7<br />

ATES – Korea Nov. 8 – 10<br />

<strong>Fluent</strong> France Nov. 9<br />

<strong>Fluent</strong> Germany Nov. 14<br />

<strong>Fluent</strong> Asia Pacific Nov. 16 – 17<br />

<strong>Fluent</strong> Italy Nov. 21<br />

Australia Nov. 21 – 22<br />

<strong>Fluent</strong> India Nov. 21 – 22<br />

<strong>Fluent</strong> in Spain Nov. 24<br />

International Oil & Gas CFD Conference Nov. 30 – Dec. 1<br />

Contact your local<br />

<strong>Fluent</strong> office for details.<br />

<strong>Fluent</strong> Worldwide<br />

Corporate Headquarters<br />

<strong>Fluent</strong> Inc.<br />

10 Cavendish Court<br />

Lebanon, NH 03766, USA<br />

Tel: 603 643 2600<br />

800 445 4454<br />

Fax: 603 643 3967<br />

Email: info@fluent.com<br />

USA Regional Offices<br />

Ann Arbor, MI 48108 Austin, TX 78746<br />

Tel: 734 213 6821 Tel: 512 306 9299<br />

Santa Clara, CA 95051 Evanston, IL 60201<br />

Tel: 408 522 8734 Tel: 847 491 0200<br />

Morgantown, WV 26505<br />

Tel: 304 598 3770<br />

European Regional Offices<br />

<strong>Fluent</strong> Benelux<br />

Wavre, Belgium<br />

Tel: 32 1045 2861<br />

Email: info@fluent.be<br />

<strong>Fluent</strong> Deutschland GmbH<br />

Darmstadt, Germany<br />

Tel: 49 6151 36440<br />

Email: info@fluent.de<br />

<strong>Fluent</strong> Europe Ltd.<br />

Sheffield, England<br />

Tel: 44 114 281 8888<br />

Email: info@fluent.co.uk<br />

<strong>Fluent</strong> France SA<br />

Montigny le Bretonneux, France<br />

Tel: 33 1 3060 9897<br />

Email: info@fluent.fr<br />

<strong>Fluent</strong> Italia<br />

Milano, Italy<br />

Tel: 39 02 8901 3378<br />

Email: info@fluent.it<br />

<strong>Fluent</strong> Sweden AB<br />

Goteborg, Sweden<br />

Tel: 46 31 771 8780<br />

Email: info@fluent.se<br />

Asian Regional Offices<br />

<strong>Fluent</strong> Asia Pacific Co., Ltd.<br />

Tokyo, Japan<br />

Tel: 81 3 5324 7301<br />

Email: info@fluent.co.jp<br />

Osaka, Japan<br />

Tel: 81 6 6359 7371<br />

<strong>Fluent</strong> Software (Shanghai) Co., Ltd.<br />

Shanghai, China<br />

Tel: 86 21 53855180<br />

Email: info_china@fluent.com<br />

<strong>Fluent</strong> India Pvt. Ltd.<br />

Pune, India<br />

Tel: 91 20 2293770<br />

Email: info@fluent.co.in<br />

Distributors<br />

ANOVA Ltd. – Turkey<br />

ATES – Korea<br />

Beijing Hi-Key Technology Corp. Ltd. – China<br />

Beijing Tianyuan – China (Icepack products only)<br />

Cavendish Instruments de Mexico, S.A. de<br />

C.V.(CIM) – Mexico, Venezuela, Argentina,<br />

Chile, Colombia<br />

CFD.HU Kft – Hungary<br />

Flowmen Technology Co., Ltd. – Taiwan<br />

Fluid Codes Ltd. – UK (sole distributor for the Middle<br />

East except for Hungary)<br />

Fluvius Pty. Ltd. – Australia & New Zealand<br />

J-ROM Ltd. – Israel<br />

Plasma Venture Ltd. – Russia (serving the Auto, Aero<br />

and Oil/Gas industries only)<br />

Process Flow Ltd. – Finland & Baltics; Russia<br />

(serving Chemical, Power and Electronics industries only)<br />

Qfinsoft – South Africa<br />

Simcon Intl (Pvt.) Ltd. – Pakistan<br />

SimTec Ltd. – Greece (serving Southeastern Europe)<br />

SMARTtech Fluidos Services & Systems, Ltd. –<br />

Brazil<br />

SymKom – Poland<br />

Techsoft Engineering s.r.o. – Czech Republic &<br />

Slovak Republic<br />

TENSOR SRL – Romania

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

Saved successfully!

Ooh no, something went wrong!