Abstract
the present paper, the effectiveness factor of porous catalytic particles is evaluated in the absence of boundary conditions symmetry over the external surface by computational fluid dynamic (CFD) techniques. The first-order kinetics of decane oxidation, already evaluated experimentally, is taken as a representative reaction. Our study arises from the fact that, in the open literature, the effectiveness factor is usually calculated considering conditions of symmetry of concentration field around particles. However, depending on the fluid dynamics of the system, such conditions are not always established and, thus, our work aims at studying for the first time the behaviour of particle catalysts with non-uniform concentration fields over the surface. In particular, the effectiveness factor of the particles in a catalytic layer is calculated in the absence of symmetry by changing several parameters (temperature, tortuosity and mean pore diameter of particle) using two different methods, named Sphere-by-Sphere (SbS) and Equisized-Volume (EV), respectively. The results of these two methods are then compared to the theoretical one obtained in the presence of spherical symmetry. As a main result, we found that, for moderately low values of Thiele modulus (<1.3 ca.), the analytical expression of the effectiveness factor obtained under spherical symmetry can be also applied in non-symmetric conditions. On the contrary, this cannot be done for higher values of Thiele modulus, for which we propose an empirical correlation of the effectiveness factor based on a corrected Thiele modulus. The efficacy of our approach is stated by the fact that pseudo-homogeneous-mode simulations of the heterogeneous system show results that match very well those obtained in heterogeneous mode, with an important reduction of calculation time and memory. The presented methodology can be also applied to n-order kinetics.
1 Introduction
The use of mesoporous particles as a support for catalyst already demonstrated to be effective in a plenty of applications, like, amongst many others, the complete oxidation of organic volatile compounds (by Pt-catalyst) (Uchisawa et al. 2008), formation of hydrogen peroxide (by Pd-catalyst) (Ogasawara et al. 1998) and Fischer-Tropsch synthesis (by Co-based catalyst) (Okabe et al. 2005).
Thanks to the possibility of fabricating mesoporous supporting particles with a characteristic size of the order of micrometre (and even smaller), they can be spread inside honeycomb catalytic converters (Uchisawa et al. 2012), whose performance strictly depends not much on the type of active metal nanoparticles deposited in it but rather on the inter- and intra-particle mass transfer resistance. Such resistance can not only falsify the actual kinetics in terms of apparent reaction order and activation energy, but also create non-uniform concentration over the external surface of the catalytic particles.
The average estimation of mass transfer resistance in different catalytic systems in homogeneous and heterogeneous mode was already studied by our research group (Caravella et al. 2012, 2016; Caravella and Sun 2016). However, those works do not consider the influence of mass transfer on the external surface concentration of reactant, whose non-uniformity makes it difficult to evaluate the effectiveness factor of the particles, which plays a crucial role in the development of whatever types of catalytic layer and, thus, in the optimal design of catalytic systems.
For this purpose, the present paper aims at providing a systematic procedure to estimate the effectiveness factor of catalyst particles in real conditions, i.e., in conditions where the concentration of the key-reactant over the external surface of the particles is not uniform, which implies absence of symmetry.
The motivation behind the present work is that the current theory on which the effectiveness factor calculation is based starts from the hypothesis of uniform concentration over the external surface of the catalyst (symmetry), which is not necessarily true in real case.
To perform this study, the effectiveness factors of the particles will be numerically evaluated by means of Computational Fluid Dynamics (CFD) simulations carried out on catalytic particles, considering the decane oxidation as a target reaction. The results from CFD will be then compared with the results from the standard theory, which allows us both to calculate the related difference as a function of the Thiele modulus and to achieve a suitable empirical functionality to estimate the effectiveness factor in the absence of symmetry.
To the best of our knowledge, there are no studies focusing on the estimation of the effectiveness factor in the absence of symmetry (see, for example, the following main works among others (Adatoz, Avci, and Keskin 2015; Delparish and Avci 2016; de Klerk, Li, and Zennaro 2013; Eigenberger and Ruppel, 2013; Hagen 2015; Kolb 2013; Liu 2014; Lehman and Larsen 2014; Nawaz 2016; Onsan and Avci 2016; Powell 2017; Sen and Avci 2014; Seader, Henley, and Roper 2016; Suarez Paris et al. 2015; Tan and Li 2013; Tezcan and Avci 2015; Xu et al. 2014). In particular, Ercan et al. (1998) examined the mutual influence of internal and external mass transfer in several zeolite catalysts for the alkylation of benzene (Ercan et al. 1998). The catalyst effectiveness factor used to quantify the internal mass transfer resistance was the canonical parameter obtained in spherical symmetry (Aris 1969; Fogler 2011; Froment and Bishoff 1979).
AL-Muftah and Abu-Reesh (2005) developed a detailed mathematical model to simulate a packed-bed reactor with immobilized enzyme considering the hydrolysis of lactose. In that work, where the reaction kinetics was taken to be of Michaelis–Menten type, conditions of spherical symmetry were considered (AL-Muftah and Abu-Reesh 2005).
Moitsheki et al. (2010) studied the non-linear problem related to the reactive diffusion inside catalysts, finding the analytical solution of the differential equation arising from the microscopic balance of the mass conservation. In that work, the mathematical symmetries of the differential equations were exploited, but uniform surface concentration is considered to reach the analytical solution (Moitsheki et al. 2010).
Lim and Dennis (2012) derived and solved a stationary non-isothermal model of reaction and diffusion in a spherical catalytic pellet with a multicomponent approach based on Maxwell-Stefan transport equations implemented in a pore network based on the Cylindrical Pore Interpolation Model (Young and Todd 2005). Also in that case, the catalyst pellets are considered to be surrounded by a uniform concentration, which limits the applications of the model to relatively small particles (high pressure drop along the reactor) (Lim and Dennis 2012).
Recently, Lenzi et al. (2017) investigated the analytical solutions of fractional diffusion equations considering three-dimensional radially-symmetric particles, allowing the external surface of the particles to be modified by time-dependent phenomena like adsorption, desorption and reaction. The impressive mathematics behind that study permitted to reach analytical solutions of the equations. However, the hypothesis of particle symmetry does not allow providing any information about the influence of possible non-uniformity of external conditions of concentration (Lenzi et al. 2017).
An interesting system in which an external asymmetry is present is represented by catalyst particles in trickle beds in which gas and liquid phases are involved in a number of applications, like hydro-desulfurization and oxidation (Tan and Smith 1980). In fact, owing to the presence of both phases over the catalyst external surface, there is a non-uniform distribution of reaction rate values, because of which the conventional expressions for the effectiveness factor cannot be applied.
On the other hand, it is not feasible to obtain an exact analytical solution for the effectiveness factor due to the equations complexity. Therefore, a numerical solution is required. To handle with such system, however, several Authors proposed different approaches, where they basically made hypotheses of uniform distribution of wet and dry parts over the catalyst surface in order to use the canonical form of the effectiveness factor in combination with the so-called particle wetting efficiency – i.e., the surface fraction covered by liquid – which was taken as a weighting factor for reaction rate (Morita and Smith 1978; Sedricks and Kenney 1973) and effectiveness factor (Hartman and Coughlin 1972; Mears 1974).
Although such a bi-phase system is quite far from the single gas-phase system considered in the present work, it is interesting that further treatments of the problem of the partial wetting given by Dudukovic (1977), and later by Dudukovic and Mills (1978), involve a modified Thiele modulus to take into account the partial wetting inside and outside the particles – in those works, the Thiele modulus was modified dividing the conventional Thiele modulus by the wetting factor.
A modification of Thiele modulus to calculate a catalyst effectiveness factor that take into account the non-symmetric conditions of concentration over the external surface is justly what we propose in this context, presenting a novel methodology that can be used by engineers, material scientists and/or other kind of end-users to design systems of catalyst particles and estimate their performances. It is furthermore remarked that, although the presented approach is applied to a first-order reaction rate, it can be also generally applied to kinetics representable by n-order expressions.
2 Description of the system
As mentioned above, the overall reference system and reaction (decane oxidation) are the same as those considered in previous works, where the reaction and mass transport performance of the catalytic layer of a honeycomb catalytic converter for decane oxidation are studied (Caravella et al. 2011, 2012). In particular, the overall multiscale approach and the system being the subject of investigation are reported in Figure 1, where the physical geometry of the catalytic layer transverse to the axial direction of the channel (Figure 1d) is reduced to the smaller geometry depicted in Figure 1e owing to the exploitation of three symmetry axes. Such a last domain is the actual domain considered for simulation (see Section 3).

Description of the overall system considered for investigation. The system subject of investigation (d) is a portion of catalytic layer deposited on the wall of the channels (b, c) of a catalytic converter (a). The actual computational domain considered for simulation (e) is obtained exploiting symmetry. Adapted and rearranged with permission from Caravella et al. (2011). Copyright 2011 American Chemical Society.
As for the reaction considered in this study, the decane oxidation occurs according to the following non-elementary overall reaction (Eq. 1):
which can be considered of the first order with a good approximation in the operating conditions accounted for in the present work (Eq. 2) (Caravella et al. 2011):
where the decane molar concentration is indicated in square brackets. The related Arrhenius-type kinetic constants are reported in Table 1 (Caravella et al. 2011).
Arrhenius parameters of the kinetic constant of the decane oxidation (Caravella et al. 2011).
Parameter | Value | Description |
---|---|---|
k c0, m3 s−1 kgcat −1 | 1.66 × 1024 | Arrhenius pre-exponential factor |
E c , kJ mol−1 | 226.783 | Activation energy |
3 Mathematical approach
3.1 System settings
The model settings of the catalytic layer are the same as those reported in ref (Caravella et al. 2012), whose sketch is recalled in Figure 2 along with the relative boundary conditions. Computational meshing, simulations and post-processing analysis were performed in the environment of the commercial software ANSYS CFX (CFX-Pre, CFX-Solver and CFD-Post), after designing the actual physical models in AutoCAD®.

Sketch of the catalytic layer investigated (Figure 1d, e) along with the corresponding boundary conditions for the two calculation approaches considered in this paper: (a) Heterogeneous approach and (b) pseudo-homogeneous one.
As for the effectiveness factor η Int, its definition is reported in Eq. (3) (see, for example (Aris 1969; Froment and Bishoff 1979; Fogler 2011):
where V Part is the particle volume, R A is the reaction rate, C A,Int is the (variable) concentration of the limiting reactant inside particle and C A,Surf is the concentration of the limiting reactant over the catalyst external surface.
Since an ensemble of particles with a reactive stream flowing along the axial direction is considered, the real conditions inside particles could be quite far from the ideal case of concentration spherical symmetry. This implies that the concentration – and, thus, the reaction rate – is generally non-uniform over the particle surface. Based on this fact, the quantity C A,Surf in Eq. 3 is calculated here as the average value of concentration over the surface (Eq. 4):
Observing Eqs. 3 and 4, it can be noted that the particle volume and the surface considered for the calculation of the averages can be chosen in several ways. In this investigation, two different approaches are used. As for the former, which will be referred to as “Sphere-by-Sphere” (SbS), the definition of the internal effectiveness factor η Int (Eq. 3) is applied by calculating the average values of Eqs. 3 and 4 over the volume and the surface of each entire particle, as indicated in Figure 3a.

Sketch of the locations chosen for calculation of the internal effectiveness factor η Int using the “Sphere-by-Sphere” approach (a), and the “Equisized-Volume” approach (b).
Differently, for the latter approach – here referred to as “Equisized-Volume” (EV) –, volume and surface are chosen to belong to two consecutive half particles (Figure 3b). The term “equisized” indicates that all volumes along the ensemble have the same length. Obviously, according to this choice, surface areas and volumes in the two methods are equivalent. The only difference is that they are shifted by the distance of a layer.
The choice for the EV approach is made basing on the consideration that, since the concentration profile due to reaction mostly develops along the axial direction, the transport phenomena of interest could be better caught within locations whose boundaries are normal to the axis. Let us remark that, in the ideal case – i.e., with spherical symmetry valid for both geometry and boundary conditions – the following expression of internal effectiveness factor holds for first-order kinetics (Eq. 5) (Aris 1969; Fogler 2011; Froment and Bishoff 1979):
where the subscript “Th” indicates “Theoretical” and the parameter ϕ is the Thiele modulus generalized with respect to geometry, defined in Eq. 6 (Aris 1969; Fogler 2011; Froment and Bishoff 1979), where r Part is the particle radius, k c,1 is the first-order kinetic constant expressed in s−1 and D A,Eff is the effective diffusivity inside particle. For the purpose of this study and without loss of generality, the diffusivity inside particle is considered to be of Knudsen-type.
In principle, there are no theoretical indications establishing that Eq. 5 can be applied also to particles where the concentration boundary conditions at the surface are non-symmetric. Therefore, a possible indication of conditions under which CFD results and analytical solution are equivalent represents new information that is very useful to simplify the simulation of complex systems of particles without loss of precision due to the non-uniformity of boundary conditions around the particles. This issue will be addressed in Section 4.2.
3.2 Simulation settings for the pseudo-homogeneous approximation
In parallel with the analysis of the intra-particle mass transport limitations, simulation should also indicate a way of simplifying the original system in order to achieve the same results in a simpler and faster way without loss of significant information. This is important, for example, in a scale-up procedure aiming at reproducing the behaviour of the original system in the frame of a more complex and multi-zone structure.
Concerning this specific case, it is not computationally feasible to consider the particle system with its heterogeneity in simulations of the overall catalytic converter. To do that, a pseudo-homogeneous approach is needed. However, the more complex the original system, the more strict the conditions under which such an approach is effectively usable can be.
Therefore, the choice of the average transport and kinetic parameters by means of which original system (heterogeneous) and pseudo-homogeneous one have to be matched can be quite hard and, in some case, simply not possible unless accepting a significant discrepancy.
With these remarks, the same simulation settings as those used in Caravella et al. (2012) are chosen here to model the particle ensembles by a pseudo-homogeneous approach, which we anticipate have shown to be effective in providing a satisfactory match.
4 Results and discussion
This section is divided in several sub-sections. In the first one, the simulation results are tested by simulating a single porous particle in spherical symmetry, for which analytical solution is available. Afterwards, the entire particle ensemble is considered, providing first an analysis of decane concentration profiles and then evaluating the internal effectiveness factor using two calculation methods. Finally, the entire system is simulated by a pseudo-homogeneous approach.
4.1 Profile analysis
In Figure 4, the local decane concentration profile in each particle is visualized. In this type of plot, our interest is to show the concentration distribution inside each particle and, thus, a local scale is used for each particle. The zones inside particles with a lower and higher concentration are respectively indicated by the blue and red colour, which a slower and a faster reaction rate respectively correspond to.

Visualization of the symmetry degree inside each particle in terms of molar concentration profiles of decane for tortuosity values of (a) 3 and (b) 10. T = 260 °C, d Pore = 5 nm.
As depicted in the plots, as the reaction rate increases with respect to the Knudsen diffusion, the concentration inside particles tends to become lower and its profiles more similar to have spherical symmetry. However, this is not true for the superficial molar concentration, which is significantly non-uniform.
As for the influence of particle tortuosity, it affects the diffusion coefficient: the higher the tortuosity, the lower the diffusion coefficient inside particle. At higher tortuosity, it is more difficult for the reactant to penetrate inside particle and, thus, the reaction tends to occur closer to the external surface.
However, in this type of plots, it is not possible to understand in what conditions the concentration symmetry inside particles is favoured, as we observe a substantial transverse concentration gradient along the axis of the considered particle ensemble. In practise, we would like to evaluate the effectiveness factor of the entire ensemble of particles, which will be assessed by evaluating the effectiveness factor of each particle of the ensemble. Nevertheless, we have a non-uniform concentration profile along the ensemble and, thus, a question arises concerning the way of evaluating the effectiveness factor of particles in our system, to which it is not possible to say a priori whether the analytical expression can be applied or not. This issue is justly dealt with in the incoming sections.
4.2 Internal effectiveness factor of particles
As shown previously, the concentration profiles inside and outside the porous catalytic particles of the here-considered ensemble are not provided with spherical symmetry. Actually, generally speaking, it is very difficult for real spherical particles to work under conditions of perfectly symmetry. Nevertheless, the catalyst effectiveness factors for spherical particles are usually calculated using the theory that considers a perfect symmetry (Alvarez-Ramirez, Ochoa-Tapia, and Valdes-Parada 2005; Burghardt and Kubaczka 1996; Mariani et al. 2009; Smith, Zahradnik, and Carberry 1975).
The goal of this section is to check if and to what extent this can be done for systems for which spherical symmetry does not hold. To do that, the internal effectiveness factor will be evaluated by CFD simulations, comparing the obtained values to those corresponding to the analytical expression (Eq. 5).
As explained in Section 3.1, in the present paper two different methods are used to evaluate the effectiveness factor inside particles η Int, hare-named as “Sphere-by-Sphere” (SbS) and “Equisized-Volume” (EV) method, respectively.
Figure 5 shows the obtained results in terms of internal effectiveness factor profiles (indicated by the respective symbols) evaluated by means of the SbS approach for a dPore = 5 nm at several temperature values and two particle tortuosity factors. The continuous lines are the values corresponding to the theory valid for spherical symmetry. It has to be remarked that, in principle, each particle has different value of η Int, and, thus, a profile is expected, whose abscissas are chosen to be located in correspondence of the centre of each particle.

Profiles of the internal effectiveness factor at several temperatures for different values of tortuosity calculated with the “sphere-by-sphere” (SbS) method for tortuosity values of (a) 3 and (b) 10. d Pore = 5 nm.
Moreover, it is necessary to indicate that the first particle is excluded from the profile evaluation, because its “special” role of particle with the highest concentration set as boundary condition does not allow a fair comparison with the other particles.
Examining the obtained results, the first thing to note is that the calculated values are practically constant at all temperatures and tortuosities considered, the differences among each other being negligible with respect to their values (<0.1%). Actually, such a similarity is quite surprising, especially if considering that the concentration conditions along the ensemble change more or less rapidly with temperature. Especially at 300 °C, where the reaction rate is considerably faster than at lower temperature values, it could be expectable a more significant scattering of calculated values. Nevertheless, no significant irregularity is observed.
As for τ Part = 10, all values of internal effectiveness factor are lower than those obtained for τ Part = 3 at the same temperature. This happens because of a larger relative difference between reaction and diffusion rate inside particles.
In fact, since Knudsen diffusion – and, thus, flux – decreases with tortuosity, less amount of reacting species passes through particles, preferring passing by. This causes a higher concentration on the surface, whose consequence is a higher reaction rate. Therefore, a smaller zone inside the catalyst particles is exploited for reaction, causing the internal effectiveness factor to be lower.
As a further remarkable result, it was found that, except for the case at 300 °C and τ Part = 10, where a discrepancy of 4.5% is observed, the internal effectiveness factor is almost coincident with that calculated from the theory exploiting symmetry. Actually, this result is quite surprising, too, because there is no apparent reason to expect that the values should be the same.
In fact, if looking back at Figure 4, the concentration is not uniform over the particle surface and there is no spherical symmetry inside it. The only symmetry that holds for particles is the one established with respect to the plane passing for the ensemble axis and normal to the open boundary.
Nevertheless, the overall result of integration provides a substantial equivalence with the analytical solution. This occurs even in zones of ensemble where decane concentration and reaction rate are extremely low. If the internal effectiveness factor is calculated according to the second method (EV), the results of Figure 6 are obtained.

Profiles of the internal effectiveness factor at several temperatures for different values of tortuosity calculated with the “Equisized-Volume” (EV) method for tortuosity values of (a) 3 and (b) 10. d Pore = 5 nm.
As shown in figure, the EV method provides results almost equivalent to those from the theory in spherical symmetry even in the case at 300 °C and τ Part = 10 (about 1.0% of discrepancy). This indicates that the differences between the SbS and EV methods could be not neglected when reaction is significantly faster than Knudsen diffusion.
To check these expectations and provide a more general view of the situation, the internal effectiveness factor (averaged along the ensemble profiles) is shown in Figure 7 as a function of the Thiele modulus (ϕ), which includes the dependence of η Int on the different factors investigated in the present paper (i.e., temperature, particle tortuosity and mean pore diameter).

Average internal effectiveness factor calculated from theory (continuous line) and for both SbS and EV method (symbols) as a function of Thiele modulus. The labels indicate the percent difference of each method from the theory in spherical symmetry.
In the figure, the values calculated by means of SbS and EV method, respectively, are shown and compared with the analytical solution (“Theory” in the plot) obtained in spherical symmetry. The labels refer to the percent difference between methods and theory.
As for the EV method, it is possible to see that the value of effectiveness factor is very close to the theoretical ones in symmetric conditions within the whole range of Thiele modulus investigated, with a maximum discrepancy of 2.1% evaluated at the highest ϕ.
The situation is different for the SbS method. In fact, below a ϕ of about 1.3, the three values (SbS, EV and Theory) are still almost coincident, with a maximum discrepancy below 1%. However, at a higher ϕ, the difference between SbS, EV and Theory becomes much more significant, reaching maximum values of 21% (SbS) and 9.7% (EV) at a ϕ of about 5.4.
An explanation for the reason why the results of the EV method are much closer to the ideal case (spherical symmetry) than those of the SbS one can be found analysing the difference between the geometries considered for calculation. When internal effectiveness factor is calculated by SbS method, the axial range involved in the evaluation is equal to the particle diameter (1 μm). However, when EV method is used, the axial range is reduced to about 0.74 μm, which is the length of each equisized volume.
Therefore, under the same conditions, the variation of concentration along the ensemble axial direction is found to be lower in the equisized volume than in the entire sphere. In other words, the concentration is “more uniform” in the second case. This difference is not evident when reaction rate is slower, because the concentration gradient is lower as well.
For the same reason, the particle volume considered for the SbS method has a concentration meanly lower than the particle volume considered for EV. Consequently, the internal effectiveness factor calculated by SbS is always lower than that calculated by the EV one under the same conditions. It is also interesting that the theoretical behaviour of the symmetric system is comprised between those of the two methods proposed.
Another important result is that, although the geometries chosen for the EV and SbS methods are different from each other, the internal effectiveness factor does not depend on the particular geometry chosen for the calculation at a low Thiele modulus (<1.3). Consequently, there is a strong indication that under these conditions Eq. 5 is applicable to any other structure (SC, BCC, random, etc). Differently, at a higher Thiele modulus (>2 ca.), these considerations cannot be generally confirmed, because the differences between EV, Theory and SbS are significant. Therefore, in this last case, the most appropriate calculation method for the effectiveness factor has to be chosen and checked carefully.
Concerning this aspect, until this point we demonstrated the existence of operating ranges where the particular calculation method chosen (Theory, EV and SbS) affect the value of the internal effectiveness factor and no indications on which “the best method” could be have been provided.
In order to provide an adequate answer to such a pending question, let us recall that the internal effectiveness factor is useful because it allows the complexity of heterogeneous catalytic particles to be by-passed by means of a homogeneous approach instead. Therefore, the choice of the appropriate effectiveness factor has to be done by using it in the pseudo-homogeneous approach and checking the difference of the obtained results from those of the heterogeneous approach. This issue will be dealt with in next section.
4.3 Pseudo-homogeneous approximation
This section is focused on the feasibility of using a pseudo-homogeneous approach that could effectively match the heterogeneous results. The simulation settings used for this analysis are those described in Section 3.2.
The comparison between the two approaches is made considering the molar concentration profile of the key-reactant (decane) for all conditions investigated (temperatures, particle tortuosity and pore diameter). This variable is chosen because it is directly proportional to the reaction rate at a fixed temperature. Therefore, by looking at this variable, it is possible to visualize the qualitative behaviour of reaction, too.
Moreover, it has to be specified that, in all comparisons, the concentration values for the heterogeneous cases are calculated as the averages over several cross sections normal to the ensemble axis including the fluid domain only, i.e., excluding the porous particles. This choice is made to make the comparison fair, because the “infinitesimal” control volume in the homogeneous case is supposed to be sufficiently larger than particles and the corresponding solution regards just the concentration in the fluid.
In this situation, the profiles inside particles are indirectly taken into account by the internal effectiveness factor only, and no further information about them can be obtained from simulations in homogeneous mode.
The first data set introduced regards conditions corresponding to low values of Thiele modulus (<0.23, Figure 8a–c), evaluated at 260 °C and τ Part = 3 by changing the particle mean pore diameter from 5 to 50 nm. This figure is divided into three sub-plots to visualize some details of heterogeneous (Figure 8b) and homogeneous (Figure 8c) concentration profiles, which cannot be analysed in the full scale figure because of being very close to each other. The profiles in homogeneous are obtained using the internal effectiveness factor calculated by means of the SbS method.

Comparison between heterogeneous (symbols with lines) and pseudo-homogeneous (continuous line) approach in terms of decane molar concentration profiles at 260 °C for a particle tortuosity of 3 (a) full-scale profiles; (b) detail of heterogeneous profiles; (c) detail of homogeneous profiles.
However, under the considered operating conditions, this choice is completely arbitrary, since all the three methods provide the same values according to the previously described Figure 7, the situation being different for a higher Thiele modulus, as will be shown later.
Considering Figure 8a, because of the low ϕ considered, the reaction is very slow and, thus, the sensitivity of the profiles on the mean pore diameter is so low that they can be hardly distinguished. Nevertheless an interesting and peculiar behaviour can be observed for a mean pore diameter of 50 nm (ϕ = 0.07).
In fact, the profile in homogeneous (Figure 8c) is lower than that at 20 nm, whereas the profile in heterogeneous (Figure 8b) is above that at 5 nm. In homogeneous case, if the mean pore diameter of particle is augmented, the only effect that is observed is the increase of Knudsen diffusivity, with a consequent larger amount of reactant consumed by reaction (lower profiles). This occurs because the pseudo-homogeneous approach is based on the spherical symmetry for particles.
However, in the heterogeneous approach, another effect is taken into account: the net mass flux of reactant passing through the particles is not null, because there is no constraint of the symmetric boundary conditions imposed on the surface. Therefore, if the Knudsen diffusion is sufficiently fast and, at the same time, the reaction is slow enough, less amount of reactant is consumed by reaction, with a consequent increase of its profile.
This is confirmed by the plot at 300 °C (Figure 9), where both heterogeneous and homogeneous profiles show the same trend. In fact, under these conditions, the reaction rate is much higher than that at 260 °C and, thus, the effect of the higher diffusion flux due to the increased diameter is completely balanced and overcome by the reaction velocity, causing the heterogeneous profile for 50 nm to be below that for 20 nm, as in homogeneous case.

Comparison between heterogeneous (symbols with lines) and pseudo-homogeneous (continuous line) approach in terms of decane molar concentration profiles at 300 °C and τ Part = 3. The data related to a d pore of 20 nm are comprised between 5 and 50 nm, very close to the latter.
The difference between the two situations can be observed better in Figure 10, where the normalised amount of decane consumed by reaction defined in Eq. 7 – actually a kind of conversion – is plotted as a function of the mean pore diameter under the same conditions as those of Figures 8 and 9.

Comparison between heterogeneous (“Het.”, empty symbols) and pseudo-homogeneous (“Hom.”, filled symbols) approach in terms of normalized decane molar concentration drop (defined in Eq. 7) at 260 and 300 °C for a particle tortuosity of 3.
Although such a difference is very small, nevertheless an important information arises from this figure. In fact, it is indicated that, under certain operating conditions, the residence time of the reactants inside catalyst can be so short that the performance of a single catalytic layer can decrease, with the overall effect for the entire catalytic converter of a productivity decrease.
This can be avoided by increasing the reaction rate, which means acting on operating conditions like temperature, or on geometrical factors like pore size and/or the catalyst distribution inside the catalytic layer. However, as anticipated before, the situation is quite different for higher values of Thiele modulus.
To investigate this range of conditions, the concentration profiles of decane are shown in Figure 11 (left-hand side) together with the profiles of the difference from the heterogeneous results (right-hand side) for increasing ϕ up to a value of 5.4. Plotting such differences is necessary in order to highlight discrepancies that otherwise would remain hidden because of the large scale used in the profile plots. Simulations in homogeneous mode are carried out by considering the three different methods used to evaluate the internal effectiveness factor of particles described above, i.e., EV, Theory and SbS. From comparing these data, we will be able to:
decide if and to what extent the pseudo-homogeneous approach is effective to replace the heterogeneous one with an acceptable approximation;
check which the best calculation method for the effectiveness factor is.

Comparison between heterogeneous (symbols) and pseudo-homogeneous (continuous lines) approach in terms of decane molar concentration profiles – left-hand side, a (1) ϕ = 1.32, b (1) ϕ = 2.96 and c (1) ϕ = 5.40 – and in terms of percent difference from the data calculated in heterogeneous case – right-hand side, a 2) ϕ = 1.32, b (2) ϕ = 2.96 and c (2) ϕ = 5.40) – at different values of Thiele modulus.
Examining the obtained results for a ϕ of 1.32 (Figure 11a.1), it is possible to observe that the differences between heterogeneous and homogeneous solutions are very small. Furthermore, the different homogeneous profiles are so close to each other that they cannot be distinguished within the whole range of the ensemble positions.
The quantification of this behaviour is provided in Figure 11a.2, where it is shown that all the three calculation methods generate concentration profiles with an error below 1.6%. In particular, the SbS method provides a maximum discrepancy (0.2%) significantly smaller than the other two. However, no clear conclusions can be drawn from this, because such differences are too small.
For this reason, also in force of the results previously shown in Figure 7 – reporting a progressive increase of the difference among the internal effectiveness factors with the Thiele modulus – the sensitivity of profiles on ϕ has been checked in Figure 11b.1, b.2, c.1 and c.2 in order to establish the most suitable calculation method. As a result, the three homogeneous profiles become more distinguishable for a progressively higher ϕ, allowing us to recognize that the SbS method provides in all cases concentration profiles much closer to the heterogeneous ones.
Even for a ϕ of about 3 (Figure 11b) – where profiles almost overlap – the maximum errors generated by the EV and Theory method are significant (11.0 and 8.6%, respectively). For the highest ϕ considered, this fact is put in evidence much more, with the errors provided by the EV and Theory method reaching 19 and 26%, respectively.
Therefore, in force of these results, we can say that, generally speaking, the best method to calculate the internal effectiveness factor for a high Thiele modulus is the SbS one, whereas the theoretical expression obtained by considering spherical symmetry can be applied only for a low Thiele modulus.
Once the best method has been identified, the question concerning the practical calculation of the internal effectiveness factor arises. In fact, the pseudo-homogeneous approach would become useless if every time a heterogeneous system should be used to evaluate the internal effectiveness factor. For this reason, we want to propose an alternative calculation method, starting from the results of Figure 7. Such a method is explained as follows.
First, let us note that the shape of the behaviour of the internal effectiveness factor calculated by the SbS method – henceforth referred to as η Int,SbS – reported in Figure 7 is qualitatively the same as that of the internal effectiveness factor calculated by the analytical expression (η Int,Th). Actually, we can say that the difference between the two profiles consists in a horizontal gap that tends to be zero for a small ϕ, progressively increasing for a higher ϕ.
From these considerations, our idea is to keep using the analytical expression represented by Eq. 5, correcting, however, the Thiele modulus in order to match the results obtained for η Int,SbS . In other words, we want to calculate ηInt,SbS by using the analytical expression of the theory, which is useful because of being simple to implement in simulation software. The procedure to do that is briefly described in Figure 12. For each ϕ, the corresponding value of η Int,SbS coming from simulations in heterogeneous is put into Eq. 8, which is solved to find the corresponding value of the modified Thiele modulus ϕ Mod.

Qualitative scheme representing the correction procedure by which to calculate the modified Thiele modulus ϕ Mod from the original one ϕ.
By performing this procedure for all values of η Int,SbS , data of ϕ Mod versus ϕ is obtained. Then, a fitting of this data set is performed with the following empirical expression (Eq. 9):
This type of polynomial form is chosen not only for its simplicity, but also because the simulation data indicate that, for a low ϕ, the values of ϕ and ϕ Mod tend to have gradually the same behaviour. This implies an expression passing through the origin and a derivative in the origin itself equal to the unity.
Moreover, the absence of the square term in Eq. 9 is because this expression with just one parameter is fully able to describe the discrepancy between the two Thiele moduli, as shown in Figure 13 and reported in Table 2.
Fitting results for Eq. (9).
Empirical correlation: | ϕ Mod = ϕ + aϕ 3 |
---|---|
a, – = | 0.009987 (≅ 0.01) |
R 2, – = | 1.0000 |
Number of data points used = | 19 |
Residual sum of squares = | 7.137 10−5 |
Range of validity | ϕ < 5.4 |
By looking at the value of the fitting parameter a, it is possible to see that its value is very close to 0.01. Therefore, without significant error, we suggest the following final expression to be used for the calculation of the modified Thiele modulus, which is valid for values of Thiele modulus lower than 5.4 (Eq. 10):
Obviously, because of the empirical nature of Eq. 10, extrapolations at values of Thiele modulus higher than those considered in the present paper has to be carried out very carefully in order to avoid significant errors in pseudo-homogeneous simulations.
However, Eq. 10 can be easily used in simulation software to correct the Thiele modulus within the range of values investigated, calculating the internal effectiveness factor in conditions of non-spherical symmetry. This allows users to avoid carrying out simulations in heterogeneous mode by performing, on the other hand, simulations in simpler pseudo-homogeneous mode without significant errors, which matches the target of the present paper.
5 Conclusions
In this investigation, the role of the mass transfer inside micrometre-sized catalytic particles was studied under conditions of non-spherical symmetry. For this purpose, the catalyst layer deposited on the internal walls of the channels of a catalytic converter is considered, taking the oxidation of decane as a representative of the first-order reactions.
To the best of our knowledge, no similar studies have been carried out yet in the open literature, as the effectiveness factor is usually calculated in conditions of symmetry around catalyst particles. Thus, our work provides new information on the behaviour of deposited catalytic particles subject to real conditions of concentration profile.
To do that, the behaviour of the ensemble of catalytic particles is simulated by computational fluid dynamics evaluating concentration profiles inside and outside particles at different values of temperature, mean pore diameter and tortuosity, considering two approaches: heterogeneous and pseudo-homogeneous one. The former takes into account the presence of the solid-gas interphase, whereas the latter treats the overall ensemble as a homogeneous system in which average values of concentration exist. The results obtained from the comparison between the two approaches provided the range of conditions in which we can use the less costly and approximated pseudo-homogeneous approach instead of the more computationally-demanding heterogeneous one.
Afterwards, the internal effectiveness factor is evaluated in a particle ensemble using two different methods, i.e., the Sphere-by-Sphere (SbS) method and the Equisized Volume (EV) one. The former consists in calculating the effectiveness factor considering each sphere of the ensemble, whereas the latter consists in calculating the effectiveness factor considering portions of catalyst volume equivalent to one sphere but sharing parts of the volumes of different spheres.
For a sufficiently low value of Thiele modulus (1.3 ca.), the so-calculated effectiveness factors were found to be constant within the whole ensemble and equivalent to the theoretical value valid for spherical symmetry of boundary conditions.
On the contrary, above this value, the SbS and EV method were found to provide values significantly different from theory – up to −21% for SbS and +9.7% for EV for a Thiele modulus of about 5.4. This fact required the choice of the best calculation method, which was made by carrying out simulations of the original heterogeneous system in homogeneous mode.
Based on our settings, the simulations in homogeneous were demonstrated to be effective in reproducing the results obtained in heterogeneous mode with a very good approximation (<0.2%). According to these simulations, the SbS method was found to be the most appropriate one to calculate the effectiveness factor up to the maximum Thiele modulus investigated, for which the results valid for spherical symmetry cannot be applied to the system investigated.
Finally, an empirical correlation was proposed to calculate the effectiveness factor under conditions of non-spherical symmetry up to a Thiele modulus of about 5.4, where it was demonstrated that the classical symmetric approach is not adequate for calculation.
List of symbols
- C
-
molar concentration [mol m−3]
- D
-
diffusion coefficient [m2 s−1]
- d Part
-
particle diameter [m]
- d Pore
-
mean pore diameter of particles [m]
- J
-
diffusive molar flux [mol m−2 s−1]
- k c
-
kinetic constant [m3 s−1 kg−1]
- k c,1
-
kinetic constant [s−1]
- M
-
molar mass [kg mol−1]
- P
-
absolute pressure [Pa]
- r
-
generic radial position from the center of the porous particles [m]
- r Part
-
radius of particle [m]
- S P
-
external surface area of a porous particles [m2]
- R i
-
volumetric reaction rate of the generic i species [mol m−3 s−1]
- T
-
temperature [K]
- V Part
-
particle volume [m3]
- x
-
molar fraction [−]
- Greek Symbols
- δ
-
inter-particle distance [m]
- ε
-
porosity [−]
- η
-
effectiveness factor [−]
- φ
-
thiele modulus [−]
- ν
-
stoichiometric coefficient [−]
- ρ
-
density [kg m−3]
- τ
-
tortuosity [−]
- Subscripts
- A
-
key-reacting species (decane)
- Eff
-
effective (for diffusivity)
- Ens
-
(particle) ensemble (considering dense particles)
- Het
-
heterogeneous
- Hom
-
(pseudo) homogeneous
- i,j
-
generic species in the diffusivity definition
- Int
-
internal (for effectiveness factor, inside the porous particles)
- Part
-
particle (for density, tortuosity and porosity)
- Surf
-
surface (of the particles)
- T
-
total (for molar concentration)
- Superscripts
- Kn
-
Knudsen (for flux and diffusivity)
- MS
-
Maxwell-Stefan (for free diffusivity)
- Acronyms
- EV
-
equisized-volume (approach)
- SbS
-
sphere-by-sphere (approach)
-
Author contributions: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.
-
Research funding: None declared.
-
Conflict of interest statement: The authors declare no conflicts of interest regarding this article.
References
Adatoz, E., A. K. Avci, and S. Keskin. 2015. “Opportunities and Challenges of MOF-Based Membranes in Gas Separations.” Separation and Purification Technology 152: 207–37, https://doi.org/10.1016/j.seppur.2015.08.020.Search in Google Scholar
Aris, R. 1969. Elementary Chemical Reactor Analysis. Englewood Cliffs: Prentice-Hall.Search in Google Scholar
AL-Muftah, A. E., and I. M. Abu-Reesh. 2005. “Effects of Simultaneous Internal and External Mass Transfer and Product Inhibition on Immobilized Enzyme-Catalyzed Reactor.” Biochemical Engineering Journal 27: 167–78, https://doi.org/10.1016/j.bej.2005.08.026.Search in Google Scholar
Alvarez-Ramirez, J., J. A. Ochoa-Tapia, and F. J. Valdes-Parada. 2005. “Dynamic Effectiveness Factor for Catalyst Particles.” Journal of Physical Chemistry B 109: 11058, https://doi.org/10.1021/jp0580266.Search in Google Scholar PubMed
Burghardt, A., and A. Kubaczka. 1996. “Generalization of the Effectiveness Factor for any Shape of a Catalyst Pellet.” Chemical Engineering and Processing 35: 65, https://doi.org/10.1016/0255-2701(95)04115-x.Search in Google Scholar
Caravella, A., S. Hara, N. Hara, A. Obuchi, and J. Uchisawa. 2012. “Three-dimensional Modeling and Simulation of a Micrometer-Sized Particle Hierarchical Structure with Macro- and Meso-Pores.” Chemical Engineering Journal 210: 363, https://doi.org/10.1016/j.cej.2012.08.101.Search in Google Scholar
Caravella, A., L. Melone, Y. Sun, A. Brunetti, E. Drioli, and G. Barbieri. 2016. “Concentration Polarization Distribution along Pd-Based Membrane Reactors: a Modelling Approach Applied to Water-Gas Shift.” International Journal of Hydrogen Energy 41: 2660, https://doi.org/10.1016/j.ijhydene.2015.12.141.Search in Google Scholar
Caravella, A., and Y. Sun. 2016. “Correct Evaluation of the Effective Concentration Polarization Influence in Membrane-Assisted Devices. Case Study: H2 Production by Water Gas Shift in Pd-Membrane Reactors.” International Journal of Hydrogen Energy 41: 11653, https://doi.org/10.1016/j.ijhydene.2015.12.068.Search in Google Scholar
Caravella, A., S. Hara, N. Hara, A. Obuchi, and J. Uchisawa. 2011. “CFD Simulation of a 3D Catalytic Layer for Decane Oxidation: Case Study of Reaction on Particle Surface.” Industrial & Engineering Chemistry Research 50: 11879, doi:https://doi.org/10.1021/ie201225r.Search in Google Scholar
Delparish, A., and A. K. Avci. 2016. “Intensified Catalytic Reactors for Fischer–Tropsch Synthesis and for Reforming of Renewable Fuels to Hydrogen and Synthesis Gas.” Fuel Processing Technology 151: 72–100, https://doi.org/10.1016/j.fuproc.2016.05.021.Search in Google Scholar
Dudukovic, M. P. 1977. “Catalyst Effectiveness Factor and Contacting Efficiency in Trickle-Bed Reactors.” AIChE Journal 23: 940–4, https://doi.org/10.1002/aic.690230624.Search in Google Scholar
Dudukovic, M. P., and P. L. Mills. 1978. “Catalyst Effectiveness Factor in Trickle-Bed Reactors.” ACS Symposium Series 65: 387.10.1021/bk-1978-0065.ch032Search in Google Scholar
Eigenberger, G., and W. Ruppel. 2013. “Catalytic Fixed-Bed Reactors.” In Ullmann’s Reaction Engineering, edited by B Elvers, 305–70. Weinheim: Wiley VCH.10.1002/14356007.b04_199.pub2Search in Google Scholar
Ercan, C., F. M. Dautzenberg, C. Y. Yeh, and H. E. Barner. 1998. “Mass-Transfer Effects in Liquid-phase Alkylation of Benzene with Zeolite Catalysts.” Industrial & Engineering Chemistry Research 37: 1724–8, https://doi.org/10.1021/ie970797b.Search in Google Scholar
Fogler, H. S. 2011. Essentials of Chemical Reaction Engineering. Boston: Pearson Education, Inc.Search in Google Scholar
Froment, G. F., and K. B. Bishoff. 1979. Chemical Reactor Analysis and Design. New York: John Wiley & Sons.Search in Google Scholar
Hagen, J. 2015. Industrial catalysis: A Practical Approach, 3rd ed. Weinheim: Wiley VCH.10.1002/9783527684625Search in Google Scholar
Hartman, M., and R. W. Coughlin. 1972. “Oxidation of SO2 in a Trickle-Bed Reactor Packed with Carbon.” Chemical Engineering and Science 27: 867, https://doi.org/10.1016/0009-2509(72)80003-4.Search in Google Scholar
Kolb, G. 2013. “Review: Microstructured Reactors for Distributed and Renewable Production of Fuels and Electrical Energy.” Chemical Engineering and Processing: Process Intensification 65: 1–44, https://doi.org/10.1016/j.cep.2012.10.015.Search in Google Scholar
de Klerk, A., Y. W. Li, and R. Zennaro. 2013. “Fischer-Tropsch Technology.” In Greener Fischer–Tropsch Processes for Fuels and Feedstocks, edited by P. M. Maitlis, and A. de Klerk, 53–79. Weinheim: Wiley VCH.10.1002/9783527656837.ch3Search in Google Scholar
Lehman, S. E., and S. C. Larsen. 2014. “Zeolite and Mesoporous Silica Nanomaterials: Greener Syntheses, Environmental Applications and Biological Toxicity.” Environmental Science Nano 1: 200–13, https://doi.org/10.1039/c4en00031e.Search in Google Scholar
Liu, H. Z. 2014. “Ammonia Synthesis Catalyst 100 Years: Practice, Enlightenment and Challenge.” Chinese Journal of Catalysis 35: 1619–40, https://doi.org/10.1016/s1872-2067(14)60118-2.Search in Google Scholar
Lim, J. Y., and J. S. Dennis. 2012. “Modeling Reaction and Diffusion in a Spherical Catalyst Pellet Using Multicomponent Flux Models.” Industrial & Engineering Chemistry Research 51: 15901–11, https://doi.org/10.1021/ie302528u.Search in Google Scholar
Lenzi, M. K., E. K. Lenzi, A. Novatski, R. Menechini Neto, and L. R. da Silva. 2017. “Fractional Diffusion Equation with Spherical Symmetry and Reactive Boundary Conditions.” Fundamenta Informaticae 151: 341–54, https://doi.org/10.3233/fi-2017-1496.Search in Google Scholar
Moitsheki, R. J., T. Hayat, M. Y. Malik, and F. M. Mahomed. 2010. “Symmetry Analysis for the Nonlinear Model of Diffusion and Reaction in Porous Catalysts.” Nonlinear Analysis: Real World Applications 11: 3031–6, https://doi.org/10.1016/j.nonrwa.2009.10.023.Search in Google Scholar
Morita, S., and J. M. Smith. 1978. “Mass Transfer and Contacting Efficiency in a Trickle-Bed Reactor.” Industrial and Engineering Chemistry Fundamentals 17: 113, https://doi.org/10.1021/i160066a008.Search in Google Scholar
Mears, D. E. 1974. “The Role of Liquid Holdup and Effective Wetting in the Performance of Trickle-Bed Reactors.” International Journal of Chemical Reactor Engineering 133: 218. 3rd.10.1021/ba-1974-0133.ch017Search in Google Scholar
Mariani, N. J., C. Mocciaro, O. M. Martinez, and G. F. Barreto. 2009. “Estimation of Effectiveness Factor for Arbitrary Particle Shape and Non-linear Kinetics.” Industrial & Engineering Chemistry Research 48: 1172, https://doi.org/10.1021/ie800578p.Search in Google Scholar
Nawaz, Z. 2016. “Heterogeneous Reactor Modeling of an Industrial Multitubular Packed-Bed Ethylene Oxide Reactor.” Chemical Engineering & Technology 39: 1845–57, https://doi.org/10.1002/ceat.201500603.Search in Google Scholar
Ogasawara, K., T. Kato, N. Okuda, T Konishi, and K. Kato. 1998. “Hydrogenation Catalyst for Production of Hydrogen Peroxide, and Method for Preparation of Same.” US Patent 5853693.Search in Google Scholar
Okabe, K., X. Li, M. Toba, and Y. Yoshimura. 2005. “Fischer-Tropsch Synthesis over Bimodal Co-ir-SiO2 Catalysts Prepared by the Alkoxide Method.” Reaction Kinetics and Catalysis Letters 86: 3, https://doi.org/10.1007/s11144-005-0288-0.Search in Google Scholar
Onsan, Z. I., and A. K. Avci. 2016. Multiphase catalytic Reactors: Theory, Design, Manufacturing, and Applications. Hoboken, NJ: Wiley.10.1002/9781119248491Search in Google Scholar
Powell, J. B. 2017. “Application of Multiphase Reaction Engineering and Process Intensification to the Challenges of Sustainable Future Energy and Chemicals.” Chemical Engineering and Science 157: 15–25, https://doi.org/10.1016/j.ces.2016.09.007.Search in Google Scholar
Seader, J. D., E. J. Henley, and D. K. Roper. 2016. Separation Process Principles With Applications Using Process Simulators, 4th ed. Hoboken, New Jersey: Wiley.Search in Google Scholar
Suarez Paris, R., L. Lopez, J. Barrientos. 2015. “Catalytic Conversion of Biomass-Derived Synthesis Gas to Fuels.” In Catalysis, Cambridge, edited by J. J. Spivey, Y. F. Han, and K. M. Dooley, 62–143. Royal Society of Chemistry.10.1039/9781782622697-00062Search in Google Scholar
Sen, I., and A. K. Avci. 2014. “Simulation of Exhaust Gas Reforming of Propane in a Heat Exchange Integrated Microchannel Reactor.” International Journal of Hydrogen Energy 39: 844–52, https://doi.org/10.1016/j.ijhydene.2013.10.148.Search in Google Scholar
Sedricks, W., and C. N. Kenney. 1973. “Partial Wetting in Trickle-Bed Reactors: The Reduction of Crotonaldehyde over a Palladium Catalyst.” Chemical Engineering and Science 28: 559, https://doi.org/10.1016/0009-2509(73)80053-3.Search in Google Scholar
Smith, T. G., J. Zahradnik, and J. J. Carberry. 1975. “Non-isothermal Inter-intraphase Effectiveness Factors for Negative Order Kinetics – CO Oxidation over Pt.” Chemical Engineering Science 30: 763, https://doi.org/10.1016/0009-2509(75)85104-9.Search in Google Scholar
Tezcan, I., and A. K. Avci. 2015. “Parametric Investigation of Oxidative Coupling of Methane in a Heat-Exchange Integrated Microchannel Reactor.” Journal of Chemical Technology & Biotechnology 90: 1827–38, https://doi.org/10.1002/jctb.4493.Search in Google Scholar
Tan, X. Y., and K. Li. 2013. “Membrane Microreactors for Catalytic Reactions.” Journal of Chemical Technology & Biotechnology 88: 1771–9, https://doi.org/10.1002/jctb.4155.Search in Google Scholar
Tan, C. S., and J. M. Smith. 1980. “Catalyst Particle Effectiveness Factor with Unsymmetrical Boundary Conditions.” Chemical Engineering and Science 35: 1601–9, https://doi.org/10.1016/0009-2509(80)80053-4.Search in Google Scholar
Uchisawa, J., T. Nanba, S. Masukawa, and A. Obuchi. 2008. “Optimization of Silica Support Pore Size for Pt Catalysts in the Complete Oxidation of Volatile Organic Compound.” Reaction Kinetics and Catalysis Letters 95: 391, https://doi.org/10.1007/s11144-008-5364-9.Search in Google Scholar
Uchisawa, J., A. Obuchi, T. Nanba, S. Hara, A. Caravella, T. Tango, T. Murakami, H. Nakagawa, T. Kogawa, and A. Abe. 2012. “Effect of Macropore Formation in Pt Catalyst Supports on the Oxidation Activity for Diesel Fuel Mist.” Industrial & Engineering Chemistry Research 51: 719, https://doi.org/10.1021/ie2018965.Search in Google Scholar
Xu, J. L., K. Li, W. Y. Shi, R. J. Li, and T. Y. Peng. 2014. “Rice-Like Brookite Titania as an Efficient Scattering Layer for Nanosized Anatase Titania Film-Based Dye-Sensitized Solar Cells.” Journal of Power Sources 260: 233–42, https://doi.org/10.1016/j.jpowsour.2014.02.092.Search in Google Scholar
Young, J. B., and B. Todd. 2005. “Modelling of Multi-Component Gas Flows in Capillaries and Porous Solids.” International Journal of Heat and Mass Transfer 48 (25−26): 5338–53, https://doi.org/10.1016/j.ijheatmasstransfer.2005.07.034.Search in Google Scholar
© 2022 the author(s), published by De Gruyter, Berlin/Boston
This work is licensed under the Creative Commons Attribution 4.0 International License.
Articles in the same Issue
- Frontmatter
- Articles
- Unified fractional indirect IMC-based hybrid dual-loop strategy for unstable and integrating type CSTRs
- Oxidative desulfurization of model and real fuel samples with natural zeolite-based catalysts: experimental design and optimization by Box–Behnken method
- Non-contact heating efficiency of flowing liquid effected by different susceptors in high-frequency induction heating system
- Gas–liquid mixing in the stirred tank equipped with semi-circular tube baffles
- Customizing continuous chemistry and catalytic conversion for carbon–carbon cross-coupling with 3dP
- Influence factor of Pr(III) recovery kinetics from rare-earth simulant wastewater by PAN microtubule hyperfiltration reactor
- NanoParticle Flow Reactor (NanoPFR): a tested model for simulating carbon nanoparticle formation in flow reactors
- Hot slag modification with mechanical stirring: heat transfer characteristics in a slag pot
- Assessment of effectiveness factor in porous catalysts under non-symmetric external conditions of concentration
- CFD simulation of gas–solid fluidized bed hydrodynamics; prediction accuracy study
Articles in the same Issue
- Frontmatter
- Articles
- Unified fractional indirect IMC-based hybrid dual-loop strategy for unstable and integrating type CSTRs
- Oxidative desulfurization of model and real fuel samples with natural zeolite-based catalysts: experimental design and optimization by Box–Behnken method
- Non-contact heating efficiency of flowing liquid effected by different susceptors in high-frequency induction heating system
- Gas–liquid mixing in the stirred tank equipped with semi-circular tube baffles
- Customizing continuous chemistry and catalytic conversion for carbon–carbon cross-coupling with 3dP
- Influence factor of Pr(III) recovery kinetics from rare-earth simulant wastewater by PAN microtubule hyperfiltration reactor
- NanoParticle Flow Reactor (NanoPFR): a tested model for simulating carbon nanoparticle formation in flow reactors
- Hot slag modification with mechanical stirring: heat transfer characteristics in a slag pot
- Assessment of effectiveness factor in porous catalysts under non-symmetric external conditions of concentration
- CFD simulation of gas–solid fluidized bed hydrodynamics; prediction accuracy study