international journal of heat and mass...

9
Numerical study of cell morphology effects on convective heat transfer in reticulated ceramics Simone Pusterla, Maurizio Barbato, Alberto Ortona , Claudio D’Angelo ICIMSI, DTI, SUPSI, Manno, Switzerland article info Article history: Received 8 November 2011 Received in revised form 26 July 2012 Accepted 7 August 2012 Available online 16 September 2012 Keywords: Ceramic foams Computational fluid dynamics Pressure drop Convective heat transfer abstract Understanding the influence of foam morphology on the heat transport mechanism is an essential task for the design engineers. The assessment of foam thermal properties was performed using experimental techniques or simulation approaches such as Finite Elements analysis and/or computational fluid dynam- ics and was, up to now, mainly focused on describing the influence of some average parameters, such as cell size and porosity. Recent numerical analysis have instead demonstrated that local cell morphological structures can strongly influence thermal conduction in ceramic foams. Therefore, in the present work, the effect of morphological characteristics, namely ligament radius, cell inclination angle and ligament tapering, on the convective heat transfer of ceramic foams were studied. The approach used is Computa- tional Fluid Dynamics (CFD) and foam geometries were schematically represented with tetrakaydecahe- dra geometries. The numerical simulations, performed with ANSYS/Fluent on different tetrakaydecahedra geometries, aimed at evaluating pressure drop and heat exchange through the foam. A heat exchanger efficiency parameter was defined and then evaluated for the different foam geometries at several air flow velocities. Results show the influence of the different morphological parameters and, in particular, that the heat exchanger efficiency of the foams decreases when increasing the air flow velocity. Ó 2012 Elsevier Ltd. All rights reserved. 1. Introduction Among the so called hybrid materials, cellular structures and foams stand for their outstanding effective properties resulting from the combination of two or more materials, or of materials and space [1]. Synthetic foams in general are widely employed in many industrial applications including: filtration, personal safety, and packaging [2], aeronautics [3], and aerospace [4]. Reticulated ceramics are highly porous open celled ceramics commonly manufactured by the replica method. Replica ceram- ics are produced in three steps: dipping a polymeric foam into a suitable ceramic slurry, de-binding and then sintering [5]. Their fabrication method is relatively economic, and large parts can be produced. Process parameters do influence the properties of the foam: by changing the slurry viscosity or by repeating the dipping phase several times, the strut (or ligament) thick- ness (i.e., porosity) can be accurately adjusted. The morphology of the foam cells and strut tapering are other two foams parameters studied in this work. They can be modified during foaming of the polymeric templates used for replica [2] or, as for cells stretching direction, it can be selected by cutting ori- ented smaller pieces from a bigger block. More recently, replica was also applied on engineered polymer scaffolds obtained by 3D printing [6]. Advanced ceramics are extensively applied as constituent mate- rials for ceramic foams [7]. In particular silicon carbide highly por- ous reticulated ceramics, thanks to their open interconnected porosity and superior thermo-mechanical properties [8,9], are used for high temperature applications where a very hot fluid flows through their porosity. These applications are: porous burners [10–13], catalytic reactors supports [14,15] and concentrated solar absorbers [16,17]. Reticulated Si–SiC foams are characterized by an interconnected network of intercommunicating cells whose edges are made of so- lid ligaments. The flowing fluid behavior through the cells, and the associated convective heat transfer between the ligaments outer surface and the fluid has been widely studied in the past years. The numerical modeling approaches used to describe these phenomena can be divided in two major classes: Local thermal equilibrium (LTE) models [18], where fluid and solid phases are in local thermal equilibrium. Local Thermal Non-Equilibrium models, which assumes that there is a finite temperature difference between the fluid and the solid phase [19]. 0017-9310/$ - see front matter Ó 2012 Elsevier Ltd. All rights reserved. http://dx.doi.org/10.1016/j.ijheatmasstransfer.2012.08.013 Corresponding author. Address: University of Applied Sciences (SUPSI), The iCIMSI Research Institute, Galleria 2, CH 6928 Manno, Switzerland. Tel.: +41 58 666 66 47; fax: +41 58 666 66 20. E-mail address: [email protected] (A. Ortona). International Journal of Heat and Mass Transfer 55 (2012) 7902–7910 Contents lists available at SciVerse ScienceDirect International Journal of Heat and Mass Transfer journal homepage: www.elsevier.com/locate/ijhmt

Upload: others

Post on 13-Apr-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: International Journal of Heat and Mass Transferrepository.supsi.ch/4867/1/pusterla-barbato-ortona-d... · 2015-07-21 · Convective heat transfer abstract Understanding the influence

International Journal of Heat and Mass Transfer 55 (2012) 7902–7910

Contents lists available at SciVerse ScienceDirect

International Journal of Heat and Mass Transfer

journal homepage: www.elsevier .com/locate / i jhmt

Numerical study of cell morphology effects on convective heat transferin reticulated ceramics

Simone Pusterla, Maurizio Barbato, Alberto Ortona ⇑, Claudio D’AngeloICIMSI, DTI, SUPSI, Manno, Switzerland

a r t i c l e i n f o a b s t r a c t

Article history:Received 8 November 2011Received in revised form 26 July 2012Accepted 7 August 2012Available online 16 September 2012

Keywords:Ceramic foamsComputational fluid dynamicsPressure dropConvective heat transfer

0017-9310/$ - see front matter � 2012 Elsevier Ltd. Ahttp://dx.doi.org/10.1016/j.ijheatmasstransfer.2012.08

⇑ Corresponding author. Address: University of AiCIMSI Research Institute, Galleria 2, CH 6928 Manno,66 47; fax: +41 58 666 66 20.

E-mail address: [email protected] (A. Orton

Understanding the influence of foam morphology on the heat transport mechanism is an essential taskfor the design engineers. The assessment of foam thermal properties was performed using experimentaltechniques or simulation approaches such as Finite Elements analysis and/or computational fluid dynam-ics and was, up to now, mainly focused on describing the influence of some average parameters, such ascell size and porosity. Recent numerical analysis have instead demonstrated that local cell morphologicalstructures can strongly influence thermal conduction in ceramic foams. Therefore, in the present work,the effect of morphological characteristics, namely ligament radius, cell inclination angle and ligamenttapering, on the convective heat transfer of ceramic foams were studied. The approach used is Computa-tional Fluid Dynamics (CFD) and foam geometries were schematically represented with tetrakaydecahe-dra geometries. The numerical simulations, performed with ANSYS/Fluent on different tetrakaydecahedrageometries, aimed at evaluating pressure drop and heat exchange through the foam. A heat exchangerefficiency parameter was defined and then evaluated for the different foam geometries at several air flowvelocities. Results show the influence of the different morphological parameters and, in particular, thatthe heat exchanger efficiency of the foams decreases when increasing the air flow velocity.

� 2012 Elsevier Ltd. All rights reserved.

1. Introduction

Among the so called hybrid materials, cellular structures andfoams stand for their outstanding effective properties resultingfrom the combination of two or more materials, or of materialsand space [1]. Synthetic foams in general are widely employed inmany industrial applications including: filtration, personal safety,and packaging [2], aeronautics [3], and aerospace [4].

Reticulated ceramics are highly porous open celled ceramicscommonly manufactured by the replica method. Replica ceram-ics are produced in three steps: dipping a polymeric foam intoa suitable ceramic slurry, de-binding and then sintering [5].Their fabrication method is relatively economic, and large partscan be produced. Process parameters do influence the propertiesof the foam: by changing the slurry viscosity or by repeatingthe dipping phase several times, the strut (or ligament) thick-ness (i.e., porosity) can be accurately adjusted. The morphologyof the foam cells and strut tapering are other two foamsparameters studied in this work. They can be modified duringfoaming of the polymeric templates used for replica [2] or, as

ll rights reserved..013

pplied Sciences (SUPSI), TheSwitzerland. Tel.: +41 58 666

a).

for cells stretching direction, it can be selected by cutting ori-ented smaller pieces from a bigger block. More recently, replicawas also applied on engineered polymer scaffolds obtained by3D printing [6].

Advanced ceramics are extensively applied as constituent mate-rials for ceramic foams [7]. In particular silicon carbide highly por-ous reticulated ceramics, thanks to their open interconnectedporosity and superior thermo-mechanical properties [8,9], are usedfor high temperature applications where a very hot fluid flowsthrough their porosity. These applications are: porous burners[10–13], catalytic reactors supports [14,15] and concentrated solarabsorbers [16,17].

Reticulated Si–SiC foams are characterized by an interconnectednetwork of intercommunicating cells whose edges are made of so-lid ligaments. The flowing fluid behavior through the cells, and theassociated convective heat transfer between the ligaments outersurface and the fluid has been widely studied in the past years.

The numerical modeling approaches used to describe thesephenomena can be divided in two major classes:

� Local thermal equilibrium (LTE) models [18], where fluid andsolid phases are in local thermal equilibrium.� Local Thermal Non-Equilibrium models, which assumes that

there is a finite temperature difference between the fluid andthe solid phase [19].

Page 2: International Journal of Heat and Mass Transferrepository.supsi.ch/4867/1/pusterla-barbato-ortona-d... · 2015-07-21 · Convective heat transfer abstract Understanding the influence

Fig. 1. Tetrakaydecahedron cell.

Nomenclature

A Area of the foam surface, m2

D Tetrakaydecahedron cell width, mh mean convective heat exchange coefficient, W/m2KH tetrakaydecahedron cell height, mk Turbulent kinetic energy, m2/s2

L Length of the major tetrakaydecahedron ligament, mLC Length of the schematic foam element, m_m mass flow rate, kg/s

p pressure, Pa_Q heat flux, WRex Cell ligament radius at the node, mRmin Cell ligament radius at its center, mt Cell ligament tapering, –

T Temperature, KTf,avg Fluid average temperature, KTf,in Fluid inlet temperature, KTf,out Fluid outlet temperature, KTs Foam surface temperature, Kv Fluid specific volume, m3/kgh Tetrakaydecahedron inclination angle, degreese rate of dissipation of turbulent energy, m2/s3

x frequency of large eddies, s�1

gth Heat exchanger efficiency, –g�th Heat exchanger efficiency, –q Fluid mass density, kg/m3

Table 1Main tetrakaydecahedron parameters [28].

Parameter Description

H = 4Lsinh Cell height [mm]

D ¼ 2 L cos hþffiffiffi2p

L � b=L Cell width [mm]

L Length of the major ligament [mm]b/L Ratio between ligament lengthsh Inclination angle (degrees)

S. Pusterla et al. / International Journal of Heat and Mass Transfer 55 (2012) 7902–7910 7903

Experimental data acquisition was diffusely utilized to charac-terize fluid flow and heat transfer in foams. The first attempt canbe traced back to Darcy’s work on porous media. More recently,Paek et al. [20] studied thermo physical properties of highly porousaluminum foams; they found that, for a given porosity value, thedecrease of cell size resulted in the increase of the surface areato volume ratio with consequent increase of flow resistance andpressure drop. According to the reported results they demon-strated that permeability is influenced by both porosity and cellsize.

Bonnet et al. [21] established simple relations between flow andmorphological parameters of metallic foams, concluding that justpore size can be sufficient for the description of flow laws in suchhighly porous materials. Hwang et al. [22] studied the heat transferand friction factor in a duct filled with aluminum foams withporosity in the range of 70–95%. These authors concluded that fric-tion factor and volumetric heat transfer coefficient increase withdecreasing sample porosity for a given Reynolds number. A similarwork was performed by Boomsma and Poulikakos [23] which com-pressed 40 PPI aluminum foams to obtain samples with differentporosities and tested them under forced convection using wateras working fluid. They observed that the structural differences,generated by compressing the specimens, did not have a noticeableeffect on permeability. Based on these observation, Boomsma et al.[24] experimentally evaluated the heat exchange performance offoams.

The present work describes how, beside porosity and cell size,foam cell parameters such as cell inclination, foam ligament shapeand cross section can influence its convective performances. As re-cently shown, these parameters can be purposely modified to de-sign and produce engineered cellular ceramics [6].

In this work, foam morphology effects on convective heat trans-fer were studied by using CFD simulations of an air flow passingthrough a foam sample which was schematically described bythree tetrakaydecahedra cells; the effects of variation of cell incli-nation angle, ligament tapering and external ligament radius wereanalyzed.

2. Foam representation

Reticulated ceramic foams can be described as a lattice of opencells with typically 12–14 pentagonal or hexagonal faces. Manyidealized geometries, with different complexity, were used in theliterature to represent and study foam thermal behavior. One ofthe simplest model is a lattice of repeating cubic cells [25]. Moreaccurate models, such as Weaire–Phelans cells [26] and the tetra-kaydecahedra [27] were also used to numerically evaluate thepressure drop and convective heat transfer in foams [25]. The

Weaire–Phelan cell is theoretically more efficient in representingthe complex foam structure compared to the tetrakaydecahedralone; however, the very limited advantages, and the increased com-plexity in building such kind of complicated geometries makes thetetrakaydecahedra cell the most widely used [25]. Due to the foam-ing process of the template foam, a better cell representation is gi-ven by the elongated tetrakaydecahedron, a polyhedron composedof 14 faces and 36 edges (Fig. 1). To unambiguously define itsgeometry, the basic geometrical parameters and correlations werecalculated by Sullivan et al. [28] (see Table 1).

In CFD studies found in the literature so far, foam structure hasbeen mainly represented as a lattice of repeating tetrakaydecahe-

Page 3: International Journal of Heat and Mass Transferrepository.supsi.ch/4867/1/pusterla-barbato-ortona-d... · 2015-07-21 · Convective heat transfer abstract Understanding the influence

Fig. 2. 3D rendering of a real foam; the red circle evidences ligament taperings. (Forinterpretation of the references to colour in this figure legend, the reader is referredto the web version of this article.)

Fig. 3. Strut geometric representation.

7904 S. Pusterla et al. / International Journal of Heat and Mass Transfer 55 (2012) 7902–7910

dra whose edges are cylinders. As shown in a 3D X-ray computedtomography rendering of real foam (Fig. 2), ligament producedby polymer replica are not cylindrical. Tapering is evident, showinghow cylinders are not adequate for their representation. Therefore,the ligament morphology was accounted in this study defining, lig-ament tapering, as the ratio between the external ligament radiusRex, at the ligaments nodal points and the radius Rmin at the liga-ment center (Fig. 3).

3. Numerical simulation

3.1. Computational mesh generation

The software MatlabV.10 [29] was used to generate a paramet-ric tetrakaydecahedra model [30]. The Matlab routines generatethe surface of the elementary cell in STL format starting from a listof input data, namely: cell size, cell inclination, ligament radius,and ligament tapering. Values ranges for these input data, suitableto generate realistic geometries for the tetrakaydecahedra cells,were defined on the basis of the results of an analysis of their ef-fects performed in [30]. The parameters range and the correspond-ing cells geometries are reported in Fig. 4. The algorithm thenconverts the geometry into a stack of 2D black (fluid) and white(foam) images, similar to Computer Tomography outputs. Theyare imported into the AvizoFire visualization software [31] forthe generation of an intermediate surface mesh of three elemen-

tary cells. The computational grid is composed by the cells surfacemesh, a fluid boundary layer composed of hexahedral elementsand an internal fluid core mesh composed of tetrahedral elements(Fig. 5). Domain meshing was performed with Altair Hypermeshsoftware [32]. To obtain an accurate description of the intricateflow path through the ligaments network, which includes imping-ing, separation and recirculation zones, the grid must be fine en-ough and grid independency of results must be guaranteed.Therefore, a grid refinement study was performed to check theindependence of simulation results from grid resolution. For oneof the schematic foam geometries, four computational grids withincreasing resolution were generated. Simulation results of heatflux through the foam surface were then compared. This analysisshowed that no further improvements were found for computa-tional meshes with a number of computational cells larger than1.6 � 106. Therefore the mesh with this number of cells was keptfor performing the further CFD analysis and worked as a referencefor the mesh generation of all other foam geometries.

3.2. Modeling equations

A tridimensional and steady air flow passing through the threecells elements was considered. It was assumed to be uncompress-ible and with constant thermal and transport properties of thefluid. In fact, the flow Mach number is, for all test cases,Ma� 0.3 and the observed temperature variations do not haveany relevant influence on the fluid properties. The flow Reynoldsnumber was calculated using the mean cell size as characteristicdimension and, on the basis of the evaluation of the critical valueof the Reynolds number for flows in porous media, the flow wasmodeled as turbulent. Therefore, the flow modeling equations in-cluded: steady state continuity, Navier–Stokes, energy, and turbu-lence model equations. Fluid density was evaluated via the idealgas law, whereas to air heat capacity, viscosity and thermal con-ductivity were given constant values. The foam was assumed asa uniform temperature solid slightly hotter than the flowing air.Convective heat transfer was evaluated whereas radiative heat ex-change was not taken into account being the fluid optically trans-parent and the foam at uniform temperature.

3.3. Computational domain and boundary conditions

The computational domain can be seen in Fig. 6: it includesthree solid cells, air inside the cells and an air duct behind the cellsexit section. Three cells, the common thickness of commercial por-ous burners [10], were considered as a sufficiently representativeelementary volume of the entire foam.

A constant velocity condition (Velocity Inlet) together with aconstant static temperature (300 K) were imposed at the computa-tional domain inlet section, i.e., at the entrance of the fluid channel(Fig. 6). In the different tests performed, the air velocity was variedto 1, 2, and 2.5 m/s. The outlet section of the computational do-main was characterized with a zero gradients condition usuallyindicated as ‘‘Pressure Outlet’’ boundary condition. There, the flowrelative static pressure was assumed to be zero (the reference pres-sure was p = 101,325 Pa). The cells were assumed to be made ofCarsik Si–SiC [33] and their surface was defined as a solid, non-slip,impermeable wall with constant and uniform temperature equal to330 K.

Symmetry boundary conditions were specified on the other fourcomputational domain external surfaces. This assumption is basedon the macroscopic characteristics of a flow entering orthogonallyinto a wide plane foam: neglecting the border effects, the averageflow can be assumed to be one dimensional with average stream-lines parallel to the inlet velocity direction. The symmetry bound-ary condition imposed to the four long external faces of the

Page 4: International Journal of Heat and Mass Transferrepository.supsi.ch/4867/1/pusterla-barbato-ortona-d... · 2015-07-21 · Convective heat transfer abstract Understanding the influence

Fig. 4. A graphical representation of the cells generated by our algorithm varying one parameter at a time.

Fig. 5. An example of the CFD mesh used in the present work. A structured mesh isused on solid walls.

S. Pusterla et al. / International Journal of Heat and Mass Transfer 55 (2012) 7902–7910 7905

computational domain implies that there the velocity is only tan-gential and that they are adiabatic.

3.4. Turbulence model

The steady state turbulent flow was solved with RANS equa-tions, k � e and k �x turbulence models were tested. Among thedifferences of these models an important one is that the formerneeds a wall treatment, whereas the latter does not. In fact, be-cause turbulent fluctuations are suppressed near the solid bound-aries and viscous effects become important, the k � e transportequations are not valid anymore there and they need to be comple-

mented by using the so called ‘‘wall functions’’. This special bound-ary treatment act as a bridge between the outer flow region andthe solid wall and avoids the need to solve explicitly the turbulentflow near the solid boundaries. The quality of the results achiev-able with the k � e models depends strongly on the proper applica-tion of the wall functions.

Other authors dealing with simulation of flows through foamshave successfully applied the SST- k–x model with low Reynoldscorrection [25]. In the present work a series of numerical testswere performed to identify the best performer between the modelsconsidered. Finally the low Reynolds nature of the flow, the possi-bility of having transitional-flow conditions for certain test cases,and the possibility of avoiding the use of wall functions, drovethe authors to select the SST- k–x. This model is in fact able ofdealing with transitional flows and, furthermore, performs verywell close to solid walls. Additionally, the SST blending offers asolution of the main k–x defect, the extreme sensitivity to turbu-lence inlet values [34].

3.5. Numerical methods

The air flow was simulated via a finite volume method approachby using the commercial CFD software ANSYS-Fluent [35]. Secondorder Upwind numerical schemes [36] were used for mass,momentum, temperature, turbulent kinetic energy, and turbulentenergy dissipation equations. The convergence of the iterativesolution process was monitored using two factors: the equationsresiduals threshold value and the stability of a monitored quantity.The former was set to 10�3 for velocity, continuity, k and x and to10�6 for the energy equation. For the latter, the computational do-main exit temperature value was chosen. Therefore, provided thatthe residuals were under the fixed threshold, the simulations werecontinued since the air exit temperature did not vary anymore for achosen number of successive iterations.

Page 5: International Journal of Heat and Mass Transferrepository.supsi.ch/4867/1/pusterla-barbato-ortona-d... · 2015-07-21 · Convective heat transfer abstract Understanding the influence

Fig. 6. The computational domain and boundary conditions.

7906 S. Pusterla et al. / International Journal of Heat and Mass Transfer 55 (2012) 7902–7910

3.6. Pressure drop evaluation

The amount of work required to pump the coolant through anheat exchanger is a critical factor for the evaluation of heat ex-change performances and can be quantified using the pressuredrop of the fluid passing through the porous medium:

Dp ¼ pinlet � poutlet ð1Þ

Since poutlet was set equal to zero in the ‘‘Pressure Outlet’’ boundary,the pressure drop across the foam corresponds to the value of pres-sure measured at the inlet of the channel, i.e., pinlet. The pressure va-lue at inlet was calculated with an area weighted average on theinlet surface for each geometry and each inlet velocity. Pressuretrends were obtained for each of the selected cell morphologicalparameters: cell inclination h, ligament tapering t, and ligament ra-dius Rex. To allow the comparison of data corresponding to foamswith different length, pressure drops were normalized with respectto the length of three cells LC (see Fig. 6):

Dp ¼ DpLC

ð2Þ

Fig. 7. CFD results, normalized pressure drop Dp vs cell inclination angle h (t = 1.01,Rex = 0.11).

3.7. Convective heat transfer and heat exchange efficiency

The heat transfer between solid and fluid can be characterizedby the mean convective heat transfer coefficient h defined as:

h ¼_Q

A ðTs � Tf ;avgÞð3Þ

where _Q is the heat flux through the solid surface, A is the heat ex-change solid surface, Ts is the surface temperature and Tfluid,avg is theaverage fluid temperature:

Tfluid;avg ¼Tf ;in þ Tf ;out

2ð4Þ

Depending on the cell morphology, the solid cell ligaments can op-pose more or less resistance to the fluid flow. Therefore, to performa weighted performance comparison of different cell morphologies,both the convective heat transfer and the pressure drop required topump the fluid through the porous medium should be taken intoaccount. In this work, the heat exchanger efficiency of the differentfoams was calculated as the ratio of the power exchanged betweenfoam and fluid, namely _Q and the pumping power:

gth ¼ _Q �_mDpq

� ��1

ð5Þ

where _m is the air mass flow rate and q is its mass density. The heatexchanger efficiency gth can be also read as the ratio of the energyexchanged per unit mass of fluid and the specific flow work:

gth ¼_Q_m

!� ðvDpÞ�1 ð6Þ

where v is the fluid specific volume.

4. Results discussion

4.1. Test cases matrix

The test cases matrix for the CFD simulations was build with the‘‘one factor at a time method’’. This lead to 21 different foam sam-ples that were then tested for three different air inlet velocitieseach. These test matrix led to a total of 63 CFD simulations. The cellmorphology factors varied are: cell inclination angle, ligamenttapering, and ligament radius. The air flow velocities applied are1, 2, 2.5 m/s; air inlet temperature and foam surface temperaturewere kept constant at 300 K and 330 K respectively. Results ofthe simulations are reported in Figs. 7, 9, 10, 12–15 and com-mented below.

4.2. Cell inclination angle

Fig. 7 shows how increasing the cell inclination angle the nor-malized pressure drop, Dp, decreases. The reason of this behavior

Page 6: International Journal of Heat and Mass Transferrepository.supsi.ch/4867/1/pusterla-barbato-ortona-d... · 2015-07-21 · Convective heat transfer abstract Understanding the influence

Fig. 8. Cell geometry evolution for the increasing cell inclination angle h (t = 1.01, Rex = 0.11). View perpendicular to the flow direction.

Fig. 9. CFD results, heat exchange performance, gth, vs cell inclination angle h(t = 1.01, Rex = 0.11).

Fig. 10. CFD results, heat exchange performance for unit foam length, g�th , vs cellinclination angle h (t = 1.01, Rex = 0.11).

S. Pusterla et al. / International Journal of Heat and Mass Transfer 55 (2012) 7902–7910 7907

is given by the reduction of tortuosity of the flow paths. This can beseen in Fig. 8 where cells with inclinations angle h of 30� and 60�appear as a compression or an extension in the flow direction ofthe undeformed reference cell with h = 45�. Therefore, whenincreasing the cell inclination angle, the struts become morealigned with the flow direction reducing, as a consequence, theaerodynamic resistance. This is an anisotropic effect that explainswhy the pressure drop increases notwithstanding the foam poros-ity reduction [30]. When the inlet air velocity is increased the pres-sure drop through the foam increases too for all h values.

The heat exchange efficiency, gth, is plotted vs the cell inclina-tion angle in Fig. 9. The trend is not monotonous and this is espe-cially evident for the case at V = 1 m/s for h < 50�. A continuous gth

increase occurs for h > 50�. It is important to observe that, contraryto the variation of cell ligaments tapering and radius, changing thecell inclination angle has influence on the foam element length LC

also. As an example we have that:

LC;h¼60� ¼ 2:26 � LC;h¼30� ð7Þ

Therefore to evaluate the foam performance excluding the depen-dence on its length, the heat exchanger efficiency was calculatedby using heat flux and pumping work per unit length. Results forthis efficiency, named g�th and reported by Fig. 10, show a continu-ous growth with cell inclination angle h.

4.3. Ligament tapering

Fig. 13 shows the trend of normalized pressure drop Dp/LC asfunction of the ligament tapering t. The three sets of data, corre-sponding to different inlet air velocities, show a fast decreases upto a tapering value around 2.5 and then tend to asymptotic values.Fig. 11, that shows graphically the evolution of channel cross sec-tion with the tapering variation, helps in finding the reason of thisbehavior. In fact, when the tapering increases from 1 to 5 the crosssection available to the fluid flow increases. Fig. 11 shows that arelevant variation of this cross section occurs passing from t = 1to t = 3, whereas the variation is much less evident when passingfrom t = 3 to t = 5. Another view of the tapering variation effect isgiven by Fig. 12 where the fluid velocity contours on a cell middleplane oriented parallel to the flow main direction are shown fort = 1 and t = 5. From this Figure, it can be seen that the flow crosssection between two adjacent cells increases with the taperingleading to an easier air passage at high t.

Similarly to what observed for cell inclination, also in the case ofligament tapering variation, an increase of fluid inlet velocity gen-erates an increase of the normalized pressure drop (see Fig. 13).

Finally, Fig. 14 shows that the increase of cells tapering is ben-eficial to the foam heat exchange efficiency. The best advantagesare for tapering variation in the first part of the t range, whereasan asymptotic behavior of the efficiency can be seen for t > 2.5.

Page 7: International Journal of Heat and Mass Transferrepository.supsi.ch/4867/1/pusterla-barbato-ortona-d... · 2015-07-21 · Convective heat transfer abstract Understanding the influence

Fig. 11. Different cross sections of the channel perpendicular to the flow direction obtained by increasing the ligament tapering factor t (h = 45�, Rex = 0.11).

Fig. 12. Fluid velocity distribution for t = 1 and t = 5 showing the development of a preferential internal flow path for t = 5 (h = 45�, Rex = 0.11).

Fig. 13. CFD results, normalized pressure drop Dp vs cell ligament tapering t(h = 45�, Rex = 0.11).

Fig. 14. CFD results, heat exchange performance, gth, vs cell ligament tapering t(h = 45�, Rex = 0.11).

7908 S. Pusterla et al. / International Journal of Heat and Mass Transfer 55 (2012) 7902–7910

4.4. Ligament radius

Fig. 15 shows the Dp behavior with ligament radius Rex varia-tion. The three data sets corresponding to the three explored air in-let velocity values show that the normalized pressure drop

increases with the external ligament radius. In fact, when Rex in-creases, the foam porosity decreases and, as a consequence, thereare larger solid obstructions to the fluid flow.

The shown pressure drop trend is relevant also for the heat ex-change performances of the foam. In fact, as shown in Fig. 16, not-withstanding the increase of heat exchange surface produced by

Page 8: International Journal of Heat and Mass Transferrepository.supsi.ch/4867/1/pusterla-barbato-ortona-d... · 2015-07-21 · Convective heat transfer abstract Understanding the influence

Fig. 15. CFD results, normalized pressure drop Dp vs cell ligament radius Rex

(h = 45�, t = 1.01).

Fig. 16. Heat exchange performance, gth, vs cell ligament radius Rex (h = 45�,t = 1.01).

S. Pusterla et al. / International Journal of Heat and Mass Transfer 55 (2012) 7902–7910 7909

the external ligament radius increase, the heat exchange efficiencydecreases. It can be noticed that the negative slope of the efficiencyis higher the larger is the air flow velocity.

5. Conclusions

In this work numerical simulations of the convective heat trans-fer in ceramic foams schematically represented by a tetrakaydeca-hedron structure were presented. The CFD simulations wereperformed with a Finite Volume Approach and the commercialsoftware ANSYS/Fluent. Different foam geometries were generatedwith the aim of evaluating the influence of morphological param-eters such as ligament tapering, external ligament radius and cellinclination angle on the foam heat exchange performance. Pressure

drop and heat exchanger efficiency for an air flow passing acrossconstant temperature hotter foams were evaluated. The main re-sults of the 63 simulations performed with the different foamgeometries can be summarized as follows:

� The normalized pressure drop decreases when ligament taper-ing and cell inclination angle increase; in fact, this geometrymodifications reduce the foam obstruction and, as a conse-quence, the resistance to the fluid flow. On the contrary,increasing ligament radius will increase the solid phase and,therefore, the obstruction to the fluid flow. This leads to highernormalized pressure drop values and to higher effective frictionfactors.� An increase of heat exchanger performance was measured

when increasing cell inclination angle and ligament tapering,whereas a decrease was noticed when increasing ligamentradius.� Ligament radius variation, i.e., effective density variation, rev-

eled to have higher relevance with respect to the other twoparameters. This difference becomes weaker for the high valuesof the ligament radius.� Increasing cell inclination angle and strut tapering, the heat

exchanger efficiency increases, whereas increasing the ligamentradius yields an efficiency reduction.� For all morphological parameters, an increase of fluid velocity

generates a pressure drop increase and a decrease of heatexchanger performances.

References

[1] M. Ashby, Hybrid materials to expand the boundaries of material-propertyspace, J. Am. Ceram. Soc. 94 (Issue Supplement s1) (2011) s3–s14.

[2] N.J. Mills, Polymer Foams Handbook, Elsevier, 2007.[3] A.R. Studart, U.T. Gonzenbach, E. Tervoort, L.J. Gauckler, Processing routes to

macroporous ceramics: a review, J. Am. Ceram. Soc. 89 (2006) 1771–1789.[4] S. Gianella, D. Gaia, A. Ortona, High temperature applications of Si-SiC cellular

ceramics, Adv. Eng. Mater. (2012), http://dx.doi.org/10.1002/adem.201200012.[5] P. Colombo, Conventional and novel processing methods for cellular ceramics,

Roy. Soc. 364 (1838) (2006) 109–124.[6] A. Ortona, S. Pusterla, C. D’Angelo, S. Gianella, Cellular ceramics produced by

rapid prototyping and replication, Mater. Lett. 80 (2012) 95–98.[7] M. Scheffler, P. Colombo, Cellular Ceramics: Structure, Manufacturing,

Properties and Applications, WILEY-VCH Verlag, 2005.[8] A. Ortona, S. Pusterla, P. Fino, F.R.A. Mach, A. Delgado, S. Biamino, Ageing of

reticulate Si–SiC foams in porous burners, Adv. Appl. Ceram. 109 (4) (2010)246–251.

[9] A. Fussel, D. Bottge, J. Adler, F. Marschallek, A. Michaelis, Cellular ceramics incombustion environments, Adv. Eng. Mater. 13 (11) (2011) 1008–1014.

[10] D. Trims, F. Durst, Combustion in porous medium – advances and applications,Combust. Sci. Technol. 121 (1996) 153–168.

[11] D. Trims, Porenbrennertechnologie – Ein berblick, Gas-Erdgas 147 (2) (2006)92–99.

[12] S. Wood, A.T. Harris, Porous burners for lean-burn applications, Progr. EnergyCombust. Sci. 34 (5) (2008) 667–684.

[13] M. Abdul Mujeebu, M.Z. Abdullah, M.Z. Abu Bakar, A.A. Mohamad, M.K.Abdullah, Applications of porous media combustion technology – a review,Appl. Energy 86 (9) (2009) 1365–1375.

[14] M.V. Twigg, J.T. Richardson, Fundamentals and applications of structuredceramic foam catalysts, Ind. Eng. Chem. Res. 46 (12) (2007) 4166–4177.

[15] J. Grosse, B. Dietrich, G. Incera Garrido, P. Habisreuther, N. Zarzalis, H. Martin,M. Kind, B. Kraushaar-Czarnetzki, Morphological characterization of ceramicsponges for applications in chemical engineering, Ind. Eng. Chem. Res. 48 (23)(2009) 10395–10401.

[16] T. Fend, High porosity materials as volumetric receivers for solar energetic,Optica Appl. 40 (2) (2010) 271–284.

[17] M. Roeb, et al., HycycleS – a project on solar and nuclear hydrogen productionby sulphur-based thermochemical cycles. in: Proceedings of the WHEC, 18thWorld Hydrogen Energy Conference, 2010.

[18] S.A. Khashan, M.A. Al-Nimr, Validation of the local thermal equilibriumassumption in forced convection of non-newtonian fluids through porouschannels, Transp. Porous Media 61 (2005) 291–305, http://dx.doi.org/10.1007/s11242-004-8305-8.

[19] P.X. Jiang, Z.P. Ren, Numerical investigation of forced convection heat transferin porous media using a thermal non-equilibrium model, Int. J. Heat Fluid Flow22 (1) (2001) 102–110.

Page 9: International Journal of Heat and Mass Transferrepository.supsi.ch/4867/1/pusterla-barbato-ortona-d... · 2015-07-21 · Convective heat transfer abstract Understanding the influence

7910 S. Pusterla et al. / International Journal of Heat and Mass Transfer 55 (2012) 7902–7910

[20] J.W. Paek, B.H. Kang, S.Y. Kim, J.M. Hyun, Effective thermal conductivity andpermeability of aluminum foam materials, Int. J. Thermophys. 21 (2000) 453–464.

[21] J.P. Bonnet, F. Topin, L. Tadrist, Flow laws in metal foams: compressibility andpore size effects, Transp. Porous Media 73 (2008) 233–254.

[22] J.J. Hwang, G.J. Hwang, R.H. Yeh, C.H. Chao, Measurement of interstitialconvective heat transfer and frictional Drag for Across Metal Foams, ASME J.Heat Mass Transfer 124 (2002) 120–129.

[23] K. Boomsma, D. Poulikakos, The effect of compression and pore size variationon the liquid flow characteristics in metal foams, ASME J. Fluids Eng. 124(2002) 263–272.

[24] K. Boomsma, D. Poulikakos, F. Zwick, Metal foams as compact highperformance heat exchangers, Mech. Mater. 35 (12) (2003) 1161–1176.

[25] Z. Wu, C. Caliot, G. Flamant, Z. Wang, Numerical simulation of convective heattransfer between air flow and ceramic foams to optimise volumetric solar airreceiver performances, Int. J. Heat Mass Transfer 54 (7) (2011) 1527–1537.

[26] D. Weaire, R. Phelan, A counter-example to kelvin’s conjecture on minimalsurfaces, Philos. Mag. Lett. 69 (2) (1994) 107–110.

[27] K. Boomsma, D. Poulikakos, On the effective thermal conductivity of a three-dimensionally structured fluid-saturated metal foam, Int. J. Heat Mass Transfer44 (4) (2001) 827–836.

[28] R.M. Sullivan, L.J. Ghosn, B.A. Lerch, A general tetrakaidecahedron model foropen-celled foams, Int. J. Solids Struct. 45 (6) (2008) 1754–1765.

[29] MathWorks. Natick, Massachusetts, USA.[30] S. Pusterla, A. Ortona, C. D’Angelo, M. Barbato, The influence of cell

morphology on the effective thermal conductivity of reticulated ceramicfoams, J. Porous Mater. 19 (2012) 307–315.

[31] VSG, Visualization Science Group, Inc. Burlington, MA USA.[32] Altair. Troy, Michigan, USA.[33] Shunk Ingenieurkeramik GmbH. Heuchelheim, Germany.[34] J. Tu, G.H. Yeoh, C. Liu, Computational Fluid Dynamics: A Practical Approach,

Elsevier, Butterworth-Heinemann, 2008.[35] Ansys Inc. Canonsburg, Pennsylvania, USA.[36] H.K. Versteeg, W. Malalasekera, An Introduction to Computational Fluid

Dynamics. The Finite Volume Method. Longman Scientific and Technical,Harlow, 1995.