Comparison of computational fluid dynamics (CFD) and pressure drop correlations in laminar flow regime for packed bed reactors and columns

Comparison of computational fluid dynamics (CFD) and pressure drop correlations in laminar flow regime for packed bed reactors and columns

    Comparison of computational fluid dynamics (CFD) and pressure drop correlations in laminar flow regime for packed bed reactors and co...

2MB Sizes 0 Downloads 74 Views

    Comparison of computational fluid dynamics (CFD) and pressure drop correlations in laminar flow regime for packed bed reactors and columns Andraˇz Pavliˇsiˇc, Rok Ceglar, Andrej Pohar, Blaˇz Likozar PII: DOI: Reference:

S0032-5910(18)30030-5 doi:10.1016/j.powtec.2018.01.029 PTEC 13115

To appear in:

Powder Technology

Received date: Revised date: Accepted date:

2 August 2017 5 January 2018 13 January 2018

Please cite this article as: Andraˇz Pavliˇsiˇc, Rok Ceglar, Andrej Pohar, Blaˇz Likozar, Comparison of computational fluid dynamics (CFD) and pressure drop correlations in laminar flow regime for packed bed reactors and columns, Powder Technology (2018), doi:10.1016/j.powtec.2018.01.029

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

ACCEPTED MANUSCRIPT

Comparison of computational fluid dynamics (CFD)

IP

SC R

for packed bed reactors and columns

T

and pressure drop correlations in laminar flow regime

NU

Andraž Pavlišič*, Rok Ceglar, Andrej Pohar, Blaž Likozar

MA

Department of Catalysis and Chemical Reaction Engineering, National Institute of Chemistry,

AC

ABSTRACT

CE P

TE

D

Hajdrihova 19, 1001 Ljubljana, Slovenia

Empirical correlation (EC) equations are still of a great designing importance for industrial plant construction. They are an indispensable modelling tool for engineers, reducing the time to find the optimal operating conditions. Nonetheless, numerical method complexity and product yield optimisation are advancing. Computational fluid dynamics (CFD) is thus nowadays applicable for optimizing chemical reactors. In contrast to EC, CFD acknowledges specific vessel geometry, where local physical and chemical phenomena, contributing to apparent catalytic turnover, prevail. Presently, EC and CFD were compared considering the pressure drop predictions within the packed bed columns for spherical, cylindrical, trilobe and quadrilobe particle packing, in order to determine the limits of EC accuracy. 52 configurations

ACCEPTED MANUSCRIPT were simulated and the estimations within EC validity range margins were in agreement with CFD (< 15%), while in extremes (non-negligible entrance and exit patterns), a 70% deviation

T

could be exceeded. Furthermore, boundary wall effects were found to be dependent on the

IP

stacked pellet shape and orientation, and did not necessarily lead to an increase of viscous

SC R

friction loss, relative to the infinitely wide systems, for the column-to-particle diameter ratios, lower than 10, which is contradictory to non-mechanistic relationship models. While the induced pressure difference within realistic fixed beds is elevated due to gas or liquid surface

NU

interaction and back-mixing, it can also be decreased by channelling/tunnelling, which is why

MA

the effective net influence should be analysed with CFD simulations, particularly in novel intensified and micronized processes, in which momentum, mass and heat transfer resistances

TE

D

are far from bulk medium continuity.

Keywords: computational fluid mechanics; modelling and optimization; laminar flow pattern;

AC

CE P

friction loss correlations; continuous packed-bed reactor; irregular particle packing

* Corresponding author. E-mail address: [email protected] (A. Pavlišič)

ACCEPTED MANUSCRIPT 1. Introduction

T

The two critical variables for the design of a packed bed (PB) reactor are usually predicted

IP

using empirical correlations (EC), the packing void fraction and the pressure drop across PB,

SC R

that were and still are a base for ongoing discussion among researchers [1–6]. Overgeneralization of typical parameters brings a high error possibility and leads to oversized reactors because of the capacity safety factor. The alternative method to handle the expected

NU

error is to conduct specific pilot plant tests with representative packing and correct those

MA

parameters [7,8]. However, these studies can be expensive and require a significant effort in investigating all influential parameters, especially if typical parameters are not constants [9].

D

In such reactors, local phenomena are dominant, and the homogeneous model using averaged

TE

EC, that PB reactor designs are based on, is often no longer valid, so adapted general models were proposed [10–13]. The emphasized non-homogeneity in real void fraction consequently

CE P

reflects in a bed-specific fluid flow profile influencing the certainty of bulk calculation with a generalized void fraction value [14–19]. The fluid flow profile is subject to more recent

AC

studies because of its complexity and difficult experimental setup.

In contrast to correlations and experimental investigations of adaptive parameters, several numerical approaches are available today to investigate the flow within particle packings, mostly based on computational fluid dynamics (CFD). New methods are possible especially because of advances in computer performance. A crucial starting point in performing a reliable CFD simulation is an accurate representation of the actual system with a geometrical model that is as realistic as possible. Although geometry can be drawn for PB reactors with a small number of particles for larger sets, a more complex methodology has to be implemented [20,21]. Three general approaches can be found throughout the literature: creating an artificial

ACCEPTED MANUSCRIPT bed using a digital packing algorithm usually based on the Monte Carlo method [22,23], mimicking a geometry based on MRI (magnetic resonance imaging) or CT (computerized

T

tomography) scan of an actual PB reactor [24], and generating a realistic simulation of reactor

IP

packing using the distinct element method (DEM) approach [6,25–31]. The last method is

SC R

especially popular in recent times because it allows creating a realistic random distribution of the fillers by simulating an actual act of gravitational packing. At the same time it is fast and

NU

inexpensive (in contrast to MRI).

MA

The objective of this study is to numerically obtain pressure drops for spherical and arbitrary particles in various PB configurations in the laminar regime with the aid of CFD. The obtained results for different geometries and flow rates are then compared with

TE

D

experimentally derived correlations. Finally, the velocity and pressure fields calculated by CFD are further on used to interpret deviation between CFD and EC pressure drop

AC

2. Methods

CE P

predictions.

2.1. Empirical equation background

Although many EC can be found in the literature for granular (packed beds) and cellular (open-cell foams) porous media, their understanding and formulation almost exclusively fall on the two historically very important equations. While Darcy’s law was very convenient for describing Newtonian fluids in the laminar regime, it failed to anticipate inertia forces when high Reynolds number were applied. Therefore an Austrian scientist P. Forchheimer (1901) in his work “Wasserbewegung durch Boden”, investigated fluid flow through porous media in

ACCEPTED MANUSCRIPT the high velocity regime [32]. He proposed a correction of Darcy’s equation by incorporating an inertia term via the kinetic energy of the fluid. The Forchheimer equation for the pressure

IP

T

drop normalised by the length of the packed bed is given as follows:

[1]

SC R

 p  2  wf  f wf . L k1 k2

NU

Where μ, ρf and wf are dynamic viscosity, density and superficial velocity of the fluid, and k1 and k2 are the Darcian and non-Darcian permeabilities, respectively. The superficial velocity

MA

is the hypothetical velocity, which the fluid would have if the channel was without packing. In contrast, the local velocity is the actual velocity of the fluid at any given position inside the

D

reactor.

TE

The most common interpretation of permeabilities in the Forchheimer equation was given by

 S V2  solid (1   )2

and k2 

3

 SV solid (1   )

,

[2]

AC

k1 

3

CE P

Ergun and Orning in 1949 [33]:

where ε and SV-solid are the open porosity and the specific surface area with respect to the solid volume, and α and β are Ergun parameters for viscous and inertial terms, respectively. By inserting Eq. 2 into Eq. 1 and substituting SV-solid by particle diameter dp, based on the geometry of a sphere dp = 6/ SV-solid, we obtain the equation commonly known as the Ergun equation [1]:

p (1   )2  w f (1   )  w f ,  AErgun  BErgun 3 2 L  dp 3 dp 2

[3]

ACCEPTED MANUSCRIPT

AErgun and BErgun being coefficients for viscous and inertial terms, respectively. The relation

* ) 6  1.75 , 8

[4]

SC R

AErgun  62  (2 *)62 150 and BErgun   6  (

IP

T

between viscous and inertial terms in Eq. 2 and 3 are expressed by:

NU

α* and β* being experimentally derived parameters that are closely related to the tortuous flow path of the fluid inside the porous media and therefore very important for understanding

MA

the differences between CFD and EC pressure drop predictions, which will be discussed in more detail in continuation. In addition, tortuosity factors which are included inside α* and β*

D

parameters, and could be determined theoretically or experimentally, are one of the reasons

TE

that Ergun equation fails to predict the pressure drop more accurately, unless spherical

CE P

particles and a column-to-particle diameter ratio higher than 10, are being used. Therefore, many alternations of the Ergun equation are being proposed to broaden the usage of EC, mainly correcting the Ergun equation by taking into consideration the influence of the

AC

confining walls and particle shape. Not going too much into details, some EC are presented in Tab. 1, which can be considered best suited for predicting pressure drops for systems used in this work according to literature and CFD results (the full list of EC can be found in the supplementary Tab. S1). Their advantages and drawbacks will be discussed later in the paper.

ACCEPTED MANUSCRIPT 2.2. Packing

T

A variety of codes can be found using DEM approach mainly due to great interest of the video

IP

games programmers. However, those codes (physics engines) are optimized for fast

SC R

calculation speed, only mimicking certain physical processes, rather than describing the real physical world. Nevertheless, due to today’s high computer power, physics engines for games and after effects in the film industry involving as much physics as possible inside the code are

NU

preferable. Therefore, they excite interest also for scientific purposes [37]. One of the

MA

programs that uses the benefits of a physics engine is software Blender (Blender Foundation), which was the tool of choice for recent scientific studies [38–40]. It is based on a large

D

collection of codes, contained in the Bullet Physics Library, used to manage the dynamics of

TE

rigid bodies by solving the Newton-Euler equations for both translational and rotational motions. Furthermore, joints and contacts are processed using relevant constraint/contact

CE P

equations. The most important advantage is that particle interactions are based on their actual

AC

surface mesh, so it is quite easy to use very complex shapes in a simulation.

Blender was used to realistically simulate the packing of particles into the reactor for different PB configurations. The reactor was represented by a simple tube with a diameter of 30 mm and a height of 50, 80 and 140 mm, depending on the size of the filler particles. Spherical and arbitrary particles (cylinders, trilobes and quadrilobes) randomly orientated in a container above the tube were allowed to fall freely under the gravitational pull into the reactor to obtain packings with a column-to-particle diameter ratio (λ) of 10, 8, 4 and 2. Additionally, for arbitrary fillers, the ratio between particle height (hp) to equivalent disk diameter (dDS) were varied between 1–4 (see Fig. 1), ending up with 52 different geometries for CFD to study. Number of particles in packed bed reactor for λ = 10, 8, 4 and 2 were approximately

ACCEPTED MANUSCRIPT 800, 400, 95 and 25, respectively.

T

2.3. Numerical setup

IP

The computational setup was designed to highly resemble the geometries used in the industry

SC R

(e.g. trilobes and quadrilobes are very common for petroleum processing). Furthermore, the shapes of filler particles were chosen on the basis of the experimental data that could be found in literature. For spheres, cylinders, trilobes and quadrilobes, beds with a height (hb) of 30 mm

NU

for λ = 10 & 8, 60 mm for λ = 4 and 120 mm for λ = 2 were prepared with DEM approach

MA

described above. Water was forced into the tubular reactor (D = 30 mm), firstly, through a preconditioning zone, where the full parabolic velocity profile was established, and secondly,

D

through the bed (see Fig. 2). Boundary conditions consisted of superficial velocity at the inlet,

TE

ambient pressure at the outlet and the no-slip condition at the reactor and particles walls. For each 52 geometries, five superficial velocities were chosen 0.01, 0.005, 0.001, 0.0005 and

CE P

0.0001 m/s, majority being in the range of laminar flow (10 < Re*; for some extreme cases Re* numbers fall into lower region of laminar-inertial flow but still far away from turbulent

AC

regime). Simulations were conducted with the steady-state incompressible solver for the laminar flow which can be found in the OpenFOAM package, with the second order scheme (the physical properties of the fluid are listed in Tab. 2). For cases that Re* exceed value of 10 turbulent modelling was applied, however, pressure drops predictions were within 1% to laminar model solver values. Solution was considered to be converged when the residuals dropped for 5 orders of magnitude for both the velocity and the pressure field.

After each geometry simulation, the tortuosity (τ) at wf = 0.01 m/s was calculated as the ratio between the total winding path and the direct path of the streamlines. Approximately 1000 streamlines were evenly distributed across the reactor inlet in order for τ to be statistically

ACCEPTED MANUSCRIPT valid.

IP

T

3. Results and Discussion

SC R

3.1. Grid independency study

NU

Another important step of the pre-processing phase is meshing the simulation domain so that the created mesh is later on used in the finite volume CFD code. The mesh generating utility

MA

snappyHexMesh, included in the open-source package OpenFOAM, was used. It generates a 3-dimensional mesh containing hexahedral cells that are automatically split to fit the

D

triangulated surface geometries of Stereolithography (STL) format. The mesh approximately

TE

conforms to the surface by iteratively refining a starting mesh and morphing the resulting split-hex mesh to the surface to obtain an unstructured mesh consisting of hexahedral (bulk)

CE P

and polyhedral cells at the walls. Previous researches indicate that this tool is an effective way of meshing the domain [38,41–44]. Because of its nature, no change of geometry posterior to

AC

the packing simulation by enlarging or shrinking particles is needed [45], so the porosity is not affected except for small bridges between the particles.

To obtain grid independent results, the mesh was refined until the steady state local velocity profile was converged (Fig. 3). Firstly, a coarse mesh was created which was gradually refined with the density profile of the nodes to be finer towards the reactor/particles walls. Secondly, bridges between particles where allowed to be additionally refined. To reach convergence, at least 10 million elements were needed and 2 additional refinements at bridges between particles. Typically the ratio between particle diameter and length of the mesh cells near the particle surface was 70 (thickness of the elements layer next to surface was  0.03

ACCEPTED MANUSCRIPT mm). At this conditions mean local velocity deviation relative to the finest mesh ( RD )

IP

T

dropped to 1% (Fig. 3).

SC R

3.2. CFD validation

NU

Although CFD has been on the rise for more than half a century, only recent decade computers allow us to solve more complex 3-D geometries. Because of that it is very

MA

important to validate CFD results. To this end, pressure drops obtained by CFD calculations were compared to experiments found in literature for the widely-studied example of PB

D

reactor packed with spheres. Due to great interest of this set up from the chemical engineer

TE

community, it can be used as the benchmark and therefore for CFD validation. Furthermore, this case is discussed in great detail, so the ability of CFD to resolve local phenomena can

CE P

contribute to a better understanding of a given problematic. On Fig. 4 CFD results are compared with EC pressure drop predictions. For EC, rather than experimental points, fits are

AC

presented on the graph. While for λ = 10 all EC give similar predictions as CFD, lower λ could lead to a large mean relative deviation (MRD; Eq. 5) with respect to EC fit:

MRD 

1 N

N

 i 1

pi ,calc  pi ,EC pi ,EC

100 .

[5]

As expected the Ergun correlation strongly deviates from CFD calculations, since it does not contain any wall corrections. However, we observe a distinguish jump in MRD at λ = 2 in comparison with MRDs at λ > 2 (Fig. 4, see supplementary Tab. S2). This goes in line with calculated tortuosities at wf = 0.01 m/s where  for λ = 2 is significantly higher than  at λ > 2

ACCEPTED MANUSCRIPT (Fig. 4; see supplementary Tab S7). Because AErgun and BErgun are kept constant in Ergun equation, although τ is significantly higher at those conditions, the Ergun equation gives

T

underestimated pressure drops. From Fig. 5 it is evident that streamlines in the case of λ = 2

IP

are more swirly and therefore need to travel a longer path (even back mixing can be detected)

SC R

in comparison with streamlines when λ = 10 was chosen. On average, the additional path that the streamlines need to travel is prolonged by 5%. The reason for that could be concluded from Fig. 6 where radial porosity distributions are presented. Voidage variation was subject to

NU

numerous studies since it is well known that it influences the pressure drop at small λ (<10)

MA

[10–12,46,46–49]. CFD results for radial porosity distributions (RPD) were compared with the experimentally derived correlation by de Klerk which confirms the resemblance to real packing [50]. RPD was found to correlate to streamline disturbances that correspond to higher

D

τ at lower λ. In theory the tortuosity for spherical particles is independent of particle size

TE

which goes in line that we calculated  to be the same for λ > 2 (see supplementary Tab S7).

CE P

However, the repetition pattern of packed spheres is disturbed at the reactor wall (Fig. 6), which leads to an uneven distribution of voidage at low λ. Therefore, a pronounced velocity

AC

gradient is established near the reactor wall (Fig. 5). This impacts the flow characteristic leading to higher inertia loses that are anticipated in Ergun equation. Furthermore, at low λ, the surface area of the reactor wall with respect to the total surface area is not negligible any more so wall friction, which is not included in the Ergun correlation, additionally increases the error. The geometry of the packing is interrupted at the wall, which causes higher nearwall porosity. This leads to the phenomenon of "flow channelling", in which case the velocity profile reaches a maximum near the wall[51]. Although channelling could further pronounce wall friction it was not detected in our case. This is in agreement with previous findings where channelling occurred only in the case of structured sphere packings [46].

ACCEPTED MANUSCRIPT The correlation proposed by Allen et al. [36], which was chosen to be presented in Fig. 4 since it gives the best overall matching (sphere + arbitrary particles) with CFD, is in good

T

agreement with CFD results for λ > 2 (MRD is lower than 6%, see Table S2-5). However, at λ

IP

= 2 the difference between Allen et al. and CFD prediction of the pressure drop remains high

SC R

as in the case of Ergun (MRD ≈ 75 %). Although Allen et al. correlation was not specifically derived for λ < 10, it gives similar results as EC (e.g. Reichelt correlation [52]) that includes the correction of the wall. However, they treat viscous and inertia forces quite differently in

NU

the laminar-inertial regime. While Reichelt correlation gives significantly more weight to

MA

viscous forces, this is not the case for Allen et al. equation. According to CFD it seems that Reichelt equation underestimate inertia forces while Allen et al. overestimate them. Nevertheless, after adding inertia to viscous forces, the pressure drop predictions are very

TE

D

much alike in both cases for λ > 2.

CE P

As expected, the lowest deviation between CFD and EC was in case where the wall effect was included inside the correlation. As example of such EC, Reichelt correlation is presented on

AC

Fig 4 (for other EC with wall correction factors see supplementary Tab. S1). In addition to correction factor M, which was firstly introduced by Mehta and Hawley [49], Reichelt tailored the Ergun equation by adding an empirically derived correction factor Bw. By applying those factors to the Ergun correlation, MRD never exceeds 15% for all λ (Fig. 4).

3.3. Pressure drops for arbitrary particles at λ > 2

As argued many times before in the literature it is highly unlikely that simple capillary models, such as the Ergun equation, would be sufficient to predict the pressure drops for arbitrary particles with great precision using universal constants. To overcome this problem,

ACCEPTED MANUSCRIPT two branches of new EC were developed, where dp was replaced by the equivalent surface volume diameter dSV = 6Vp /Ap . First, the already existing correlations for spherical particles

T

were simply tailored by fitting new constants for either each respective geometry or groups of

IP

geometries with similar shape [11]. Second, correction factors were added to EC and so they

SC R

become shape depended [35]. Alternatively, EC fitted for spherical particles by replacing dp with dSV can be used, however, high errors are expected. Below, one example per each approach is presented and compared with CFD results chosen on the basis of the highest

NU

relevance with simulated particle geometry according to literature (for comparison between CFD and the rest of EC see Tab. S2-S6). Since it was already shown in the case of spherical

MA

particles that a high error is induced at λ = 2, unless EC with wall correction is used, overall MRDs on Fig. 7–9 are referring to all λ > 2, however, all data points calculated with CFD are

TE

D

presented on the graphs. Due to the high differences between EC and CFD at λ = 2, special care is needed to interpret the deviation between them, which will be discussed in next

CE P

section.

AC

Latter solution for the prediction of pressure drops in the case of arbitrary particles by EC is shown on Fig. 7, where CFD results are compared with the Ergun correlation. Regarding to experimental data from literature, Ergun equation should give under predicted pressure drops for cylinders and polylobed extrudates (PE), which is in agreement with CFD results [35,36,53]. As expected, overall MRD is rising with the complexity of the geometry, since parasitic drag is dependant of the shape of the particle (see Fig. 7). Naturally, this approach gives the highest error and should be avoided if possible. According to CFD calculations, EC predicted pressure drops for cylinders would be about 30% of the real value and with PE error jumps to ~50% for λ > 2.

ACCEPTED MANUSCRIPT The correlation that was developed to cover the particle shape effect is presented on Fig. 8 along with the CFD results. Nemec&Levec enhanced the Ergun equation by adding to it the

T

sphericity factor (Φ) [35]. With this correction, the overall MRD for individual geometry

IP

(cylinder, trilobe, quadrilobe) is ~13% (see Fig. 8) which is close to the estimated accuracy of

SC R

the correlation (authors stated that the correlation is able to predict the pressure drop within 10%). It should be noted that Φ exponent is different for cylinders and PE, however, after the linearization of Nemec&Levec correlation, both can be presented on the same graph. Even

NU

though Φ brings EC and CFD simulated pressure drop predictions closer together, deviations (EC vs. CFD) vary with the hp/dDS ratio (ζ, see Tab. S2-5), but not between different geometry

MA

groups. This suggest that Φ alone is not sufficient enough to smooth the effect of geometry, so another aspect needs to be considered – wetted surface, which will be discussed in next

TE

D

paragraph.

CE P

The third option in dealing with pressure drops for arbitrary particles by using correlations is to fit constants for each specific geometry group (Fig. 9). Allen et al. in their work proposed

AC

an equation where rather than dSV, the total particle volume ∑Vp to surface ∑Sp ratio is used [36]. They stated that ∑Vp / ∑Sp ratio should be understood to refer to the duct size rather than an equivalent particle size (dSV = 6∑Vp / ∑Sp), as used by Ergun. The correlation highly resembles the Kozeny–Carman equation, but they broaden the usage by separately fitting constants not only for spheres but also cylinders, cubes, acorns and crushed rocks [2]. Surprisingly, Allen correlation gave the least deviation with respect to CFD for PE, although according to literature that should be Nemec&Levec. Overall MRD for cylinder beds is ~15 % and ~9 % for trilobe and quadrilobe beds (see Fig. 8 and 9), which is significantly lower for the latter two geometries than in the case of Nemec&Levec correlation. It is worth knowing that Nemec&Levec specifically fit their correlation for PE while constants for cylinders were

ACCEPTED MANUSCRIPT used in the case of Allen equation (since it is the closest geometry to PE). Therefore, we expect that Nemec&Levec would outperform the Allen correlation and the latter would be

T

true if we would follow the Allen equation to the letter and use the theoretical volume and

IP

surface area of the particles. However, geometry parameters were taken from the obtained

SC R

meshes. While the particles volume was always close to the theoretical value (~99%), their surface area could differ from the theoretical for more than 20% in some cases. One may argue that this is the error from the meshing procedure, but it should be noted that in the case

NU

of spheres, the surface area was always > 95% for all λ in comparison with the theoretical

MA

value. Hence, we cannot rule out that minor error is introduced during meshing, however, in the case of the arbitrary particles, the majority of the deviation from the theoretical surface area is contributed from another factor. Unlike spheres, arbitrary particles may consist of a

TE

D

certain number of flat surfaces. During the packing, those surfaces could overlap and therefore give a lower wetted surface than theoretical values. This affects the skin friction

CE P

drag and consequently the pressure drop of PB reactor. Dolejš&Machač proposed a correlation which included a correction factor for the wetted surface (ω), but ω was obtained

AC

by fitting the experimental data for a specific geometry and was not measured directly [54]. Unless investigated PB reactor falls on the particle geometry that was already covered in their work, additional experiments of the pressure drop for each particle shape needs to be performed individually. This is very time consuming, and moreover, much simpler correlations could be used if experiments were carried out and fit constants to conform specific particle geometry without any radical decrease of accuracy in comparison with Dolejš&Machač.

Going back to the Allen correlation, if surface areas obtained from meshing procedure are used, not only does the overall MRD outperformed Nemec&Levec correlation, but also

ACCEPTED MANUSCRIPT significantly reduces ζ impact on the pressure drop deviation between CFD and EC (Fig. 10, see Tab. S2-5). This indicates that ω is more relevant to the pressure drop than Φ in the

T

laminar regime. However, it is difficult to properly measure the wetted surface directly, so for

IP

now Φ correction factor is the best choice for predicting the pressure drop with EC using the

SC R

least amount of effort. Nemec&Levec equation is closely followed by Eisfeld&Schnitzlein correlation [55], which gives slightly better results at higher λ (see Tab. S2-6), but CFD data

NU

are much more scattered around the correlation curve (see Fig. 8 and Fig. S1).

On Fig. 11 are presented typical representation schemes of the results obtained by CFD

MA

simulation for quadrilobes at wf = 0.01 m/s, λ = 10 and ζ = 4. The velocity field is highly inhomogeneous along the radial axis of the packed bed reactors (to some degree this also true

TE

D

for pressure field), which put homogeneity models for solving mass transport with chemical reactions within such reactors into question. Furthermore, from this point of view it is

CE P

surprisingly that EC for pressure drop predictions offer so much accuracy. However, as would be explained in the next section, this is true only for the reactors that contain enough amounts

AC

of filler particles which consequently statistically averaged the local phenomena effects.

3.4. Pressure drop for arbitrary particles at λ = 2

As mentioned before, EC pressure drop deviation for arbitrary particles increases drastically in comparison with CFD results when λ < 4. While spheres were subject to numerous studies at low λ ratios, arbitrary particles were not given as much attention. Usually findings from spherical particles were taken and applied to other geometries. What does not make the relation between spherical and non-spherical particles straightforward is that spheres cannot

ACCEPTED MANUSCRIPT influence the pressure drop with their orientation. As shown by Pahl in his early work in 1975, the opposite stands for arbitrary particles [56]. While it was known before that the

T

porosity function inside the Ergun equation will be inadequate for non-spherical particles,

IP

engineers were still puzzled, because the pressure drop data were much more scattered than

SC R

anticipated. Pahl attributed this effect to the particle orientation by investigating cylinders and their impact on the pressure drop while varying ζ and the packing procedure. His assumption was later confirmed by several studies [57,58]. In line with previous findings, CFD results of

NU

the present work also show a significant impact on the pressure drop due to particle orientation. On Fig. 12, mean particle orientations (MPO) of quadrilobes for all combinations

o

MA

of λ and ζ with corresponding MRDs (CFD vs. Allen) are presented (similar results stand also for other two arbitrary geometries; 90 means that the particles are parallel with their largest

TE

D

dimension in the direction of the reactor axis). As expected, the orientations at λ = 2 significantly deviate from the orientations at higher λ, since at these conditions the driving

CE P

force of the reactor wall to orientate the particles vertically is the largest. While there is a distinctive trend of raising the percentage of particles being oriented vertically with raising ζ

AC

at λ = 2, it gets less pronounced at lower λ and almost reversed at λ = 10. At high λ, the o

dependency of ζ on particle orientation is rather poor, with MPO around 25 . This is in agreement with the Vollmari et al. experiment where they measured MPO for cylinders at ζ = o 3.5 and λ = 15 to be 24 [6]. Next, tortuosity was calculated for all λ at ζ = 4 and wf = 0.01 m/s

(these conditions were chosen since the impact of ζ on particle orientation was the largest). Surprisingly, τ was independent of the shape group and therefore, average τ of all three arbitrary particles are labelled on Fig. 12 (for τ values of individual geometries see Tab. S7). Tortuosities for λ > 2 were fairly close to each other, however, at λ = 2, τ is significantly lower (it drops from ~1.2 to ~1.1), most likely due to preferential orientation (MPO at this o

conditions was 76 for ζ = 4). This may explain why there is such a big difference in the

ACCEPTED MANUSCRIPT pressure drop predictions between CFD and EC at low λ. In addition, tortuosities for all ζ at λ = 2 and wf = 0.01 m/s were calculated. Again, τ was independent of the particle shape group.

T

In addition, for all ζ > 1 τ seems to be also independent of ζ. All in all, no distinguishing trend

IP

could be found between τ and particle shape. On the other hand, MPO seems to have

SC R

influence on the τ which has a significant impact on the pressure drop, since every deviation of those two parameters from values obtained for an infinite wide channel (we can

NU

approximate this for λ = 10) leads to higher MRD.

Lastly, the influence of confining walls on the pressure drop for arbitrary fillers was

MA

investigated by plotting modified friction factors vs. λ, as computed from CFD simulations for Rep=2 (Figs. 13–14). Results suggest that high ζ (=4) at low λ (=2) could lead to a reduction

TE

D

of the pressure drop (this effect is especially evident for cylindrical geometry). This is contradictory to common belief since the wall effect should increase the pressure drop at these

CE P

conditions. Low tortuosity and high MPO may be the reason that suppresses the effect of the confining wall. However, at low ζ (=1) and λ (=2) this phenomenon disappears and the system

AC

behaves as normally predicted in the literature (Fig.14).

4. Conclusions

To conclude, the pressure drops inside PB reactor for spherical and arbitrary particles (cylinders, trilobes and quadrilobes) were investigated by the aid of CFD. Results obtained by CFD were compared with the most common EC and a general agreement exists between them when the packed bed characteristic parameters were inside EC validity. For some extreme

ACCEPTED MANUSCRIPT cases, such as λ = 2, MRD raised over 50% which makes EC quite impractical. By means of tortuosity and particle orientation it was shown that at these conditions, EC are inadequate to

T

fully balance those two parameters. Their values are mostly associated with form drag and

IP

therefore at lower Re number their impact on the pressure drop is less pronounced. In

SC R

addition, MPO and τ seem to be constants for infinitely wide reactors, which in practice means for all λ > 10. Therefore, no special care is needed for EC. Fitting constants to reflect the specific geometry is sufficient that EC give solid predictions. Furthermore, the most

NU

relevant variable for predicting the pressure drop in the laminar regime was found to be the

MA

wetted surface. Although skin friction drag dominates at these conditions for higher Re numbers (but still in laminar regime) the pressure drop contribution from form drag is not negligible anymore. The latter is evident when comparison is made between Allen [36] and

TE

D

Nemec&Levec [35] EC, where the Allen correlation outperforms Nemec&Levec at lower Re numbers, but vice versa is true for higher Re numbers. This indicates that Φ inside

CE P

Nemec&Levec correlation is more related to form drag and therefore more appropriate for use at higher Re numbers. All in all, it seems that ω, τ and MPO are indispensable parameters for

AC

pressure drop predictions in case of arbitrary particles if we want to use EC at λ < 4. Unfortunately, all three parameters are rather hard to obtain by experiments so Nemec&Levec correlation remains the best choice for arbitrary particle investigated in this paper. On the other hand, CFD has no problem evaluating those bulk parameters. Even more, it provides a full inside view on the local phenomena which can greatly contribute to a better understanding of PB reactors. Lastly, the described tasks are a first step in building a stable framework for the virtual modelling of PB reactors and will continue in coupling validated fluid flow models with other chemical transport phenomena and reaction kinetics. This will provide a good basis for the unification, gradual supplementation and expansion of the empirical methods for the determination of PB reactor design parameters.

ACCEPTED MANUSCRIPT

shape defining parameter [/]

AErgun

Ergun coefficient for viscous energy losses [/]

Ap

particle surface area [m2]

BErgun

Ergun coefficient for kinetic energy losses [/]

BW

wall correction term [/]

c

shape defining parameter [/]

D

column diameter [m]

dDS

equivalent disk diameter [m]

dp

particle diameter [m]

dSV

equivalent surface volume diameter [m]

f*

modified friction factor [/]

fAllen

Allen modified friction factor [m-1]

fN&L

Nemec&Levec modified friction factor [/]

hp

particle height [m]

L

SC R

NU

MA

D

TE

CE P

AC

hb

IP

ai

T

Nomenclature

bed height [m] length of the packed bed reactor [m]

M

modification factor [/]

N

number of samples [/]

r

distance from reactor axis [m]

R

reactor radius [m]

Re*

Ergun modified Reyndols number [/]

ReAllen

Allen modified Reyndols number [m-1]

ReDuct

modified Reyndols number for duct [/]

ACCEPTED MANUSCRIPT Nemec&Levec modified Reyndols number [/]

Rep

modified Reyndols number for particles [/]

SV-solid

specific surface area with respect to the solid volume [m2]

T

temperature [K]

Vp

particle volume [m3]

wf

superficial velocity [m s-1]

NU

Greek symbols

SC R

IP

T

ReN&L

Ergun coefficient for viscous term [/]



Ergun coefficient for inertial term [/]



column-to-particle diameter ratio [/]



tortuosity [/]



wetted surface [m2]



dynamic viscosity [Pa s]

*

Ergun coefficient for viscous term [/]

*

Ergun coefficient for inertial term [/]

pi,calc

D TE

CE P

AC

p

MA



pressure drop [Pa] pressure drop predicted by CFD calculation [Pa]

pi,EC

pressure drop predicted by EC [Pa]

ε

porosity [/]

ζ

hp/dDS ratio [/]

ρf

density of the fluid [kg m-3]

Φ

sphericity [/]

ΦAllen

packing dimension [m]

ACCEPTED MANUSCRIPT Abbreviations computational fluid dynamics

DEM

distinct element method

EC

empirical correlations

MPO

mean particle orientations

MRD

mean relative deviation

PB

packed bed

RPD

radial porosity distributions

MA

NU

SC R

IP

T

CFD

Subscripts fluid

i

index

p

particle

b

bed

V-solid

solid volume

AC

CE P

TE

D

f

Acknowledgements

The authors gratefully acknowledge the European Commission, as the herein-presented research work was partially established within the MefCO2 project. This project has received funding through a Sustainable Process Industry through Resources an Energy Efficiency (SPIRE) call under the European Union’s Horizon 2020 research and innovation programme (grant agreement n° 637016). The authors also gratefully acknowledge the Slovenian Research Agency (ARRS) for funding the program P2–0152 and project J2–7319.

ACCEPTED MANUSCRIPT References

S. Ergun, Fluid flow through packed columns, Chem. Eng. Prog. 48 (1952) 89–94.

T

[1]

P.. Carman, Fluid flow through granular beds, Trans. Inst. Chem. Eng. 15 (1937) 150–

SC R

[2]

IP

doi:citeulike-article-id:7797897.

166.

http://www.citeulike.org/group/1336/article/711017\npapers3://publication/doi/10.1234

J. Kozeny, Über kapillare Leitung des Wassers im Boden, Akad. Wiss.Wien. 136

MA

[3]

NU

/12345678.

(1927) 271–306.

S.P. Burke, W.B. Plummer, Gas Flow through Packed Columns 1, Ind. Eng. Chem. 20

D

[4]

[5]

F.C. Blake, The resistance of packing to fluid flow, Trans. Am. Inst. Chem. Eng. 14

CE P

(1922) 415–421. [6]

TE

(1928) 1196–1200. doi:10.1021/ie50227a025.

K. Vollmari, T. Oschmann, S. Wirtz, H. Kruggel-Emden, Pressure drop investigations

AC

in packings of arbitrary shaped particles, Powder Technol. 271 (2015) 109–124. doi:10.1016/j.powtec.2014.11.001. [7]

M. Leva, Fluidisation, McGraw-Hill Book Company, New York, NY, 1959.

[8]

I.F. Macdonald, M.S. El-Sayed, K. Mow, F. a L. Dullien, Flow through Porous Mediathe Ergun Equation Revisited, Ind. Eng. Chem. Fundam. 18 (1979) 199–208. doi:10.1021/i160071a001.

[9]

R.E. Hicks, Pressure Drop in Packed Beds of Spheres, Ind. Eng. Chem. Fundam. 9 (1970) 500–502. doi:10.1021/i160035a032.

[10]

N.-S. Cheng, Wall effect on pressure drop in packed beds, Powder Technol. 210 (2011) 261–266. doi:10.1016/j.powtec.2011.03.026.

ACCEPTED MANUSCRIPT [11]

B. Eisfeld, K. Schnitzlein, The influence of confining walls on the pressure drop in packed beds, Chem. Eng. Sci. 56 (2001) 4321–4329. doi:10.1016/S0009-

R. Di Felice, L.G. Gibilaro, Wall effects for the pressure drop in fixed beds, 2004.

IP

[12]

T

2509(00)00533-9.

[13]

SC R

doi:10.1016/j.ces.2004.03.030.

M. Winterberg, E. Tsotsas, Impact of tube-to-particle-diameter ratio on pressure drop in packed beds, AIChE J. 46 (2000) 1084–1088. doi:10.1002/aic.690460519. M. Giese, K. Rottschafer, D. Vortmeyer, Measured and modeled superficial flow

NU

[14]

doi:10.1002/aic.690440225. [15]

MA

profiles in packed beds with liquid flow, AIChE J. 44 (1998) 484–490.

L.H.S. Roblee, R.M. Baird, J.W. Tierney, Radial porosity variations in packed beds,

[16]

TE

D

AIChE J. 4 (1958) 460–464. doi:10.1002/aic.690040415. O. Bey, G. Eigenberger, Fluid flow through catalyst filled tubes, Chem. Eng. Sci. 52

[17]

CE P

(1997) 1365–1376. doi:10.1016/S0009-2509(96)00509-X. C.E. Schwartz, J.M. Smith, Flow Distribution in Packed Beds, Ind. Eng. Chem. 45

[18]

AC

(1953) 1209–1218. doi:10.1021/ie50522a025. F. Benyahia, K.E. O’Neill, Enhanced Voidage Correlations for Packed Beds of Various Particle Shapes and Sizes, Part. Sci. Technol. 23 (2005) 169–177. doi:10.1080/02726350590922242. [19]

E.A. Foumeny, F. Benyahia, Predictive characterization of mean voidage in packed beds, Heat Recover. Syst. CHP. 11 (1991) 127–130. doi:10.1016/0890-4332(91)90126O.

[20]

M. Nijemeisland, Verification Studies of Computational Fluid Dynamics in Fixed Bed Heat Transfer, 2000.

[21]

M. Nijemeisland, A.G. Dixon, Comparison of CFD simulations to experiment for

ACCEPTED MANUSCRIPT convective heat transfer in a gas-solid fixed bed, Chem. Eng. J. 82 (2001) 231–246. doi:10.1016/S1385-8947(00)00360-0. M.J. Baker, G.R. Tabor, Computational analysis of transitional air flow through

T

[22]

IP

packed columns of spheres using the finite volume technique, Comput. Chem. Eng. 34

[23]

SC R

(2010) 878–885. doi:10.1016/j.compchemeng.2009.10.013.

R. Caulkin, A. Ahmad, M. Fairweather, X. Jia, R.A. Williams, An investigation of sphere packed shell-side columns using a digital packing algorithm, Comput. Chem.

M.J. Baker, P.G. Young, G.R. Tabor, Image based meshing of packed beds of

MA

[24]

NU

Eng. 31 (2007) 1715–1724. doi:10.1016/j.compchemeng.2007.03.014.

cylinders at low aspect ratios using 3d MRI coupled with computational fluid dynamics, Comput. Chem. Eng. 35 (2011) 1969–1977.

[25]

TE

D

doi:10.1016/j.compchemeng.2011.03.017. T. Eppinger, K. Seidler, M. Kraume, DEM-CFD simulations of fixed bed reactors with

CE P

small tube to particle diameter ratios, Chem. Eng. J. 166 (2011) 324–331. doi:10.1016/j.cej.2010.10.053. M. Kuroki, S. Ookawara, K. Ogawa, A High-Fidelity CFD Model of Methane Steam

AC

[26]

Reforming in a Packed Bed Reactor, J. Chem. Eng. JAPAN. 42 (2009) s73–s78. doi:10.1252/jcej.08we256. [27]

M. V. Tabib, S.T. Johansen, S. Amini, A 3D CFD-DEM Methodology for Simulating Industrial Scale Packed Bed Chemical Looping Combustion Reactors, Ind. Eng. Chem. Res. 52 (2013) 12041–12058. doi:10.1021/ie302028s.

[28]

T. Tsory, N. Ben-Jacob, T. Brosh, A. Levy, Thermal DEM–CFD modeling and simulation of heat transfer through packed bed, Powder Technol. 244 (2013) 52–60. doi:10.1016/j.powtec.2013.04.013.

[29]

G.D. Wehinger, T. Eppinger, M. Kraume, Detailed numerical simulations of catalytic

ACCEPTED MANUSCRIPT fixed-bed reactors: Heterogeneous dry reforming of methane, Chem. Eng. Sci. 122 (2015) 197–209. doi:10.1016/j.ces.2014.09.007. N. Zobel, T. Eppinger, F. Behrendt, M. Kraume, Influence of the wall structure on the

T

[30]

IP

void fraction distribution in packed beds, Chem. Eng. Sci. 71 (2012) 212–219.

SC R

doi:10.1016/j.ces.2011.12.029. [31]

C.A. de de Moura Teixeira, Integrated Master in Chemical Engineering, 2013.

[32]

P. Forchheimer, Wasserbewegung Durch Boden, Zeit. Ver. Deutsch. Ing. 45,

S. Ergun, A.A. Orning, Fluid Flow through Randomly Packed Columns and Fluidized

MA

[33]

NU

Zeitschrift Des Vereins Dtsch. Ingenieure. (1901) 1781–1788.

Beds, Ind. Eng. Chem. 41 (1949) 1179–1184. doi:10.1021/ie50474a011. [34]

W. Reichelt, Zur Berechnung des Druckverlustes einphasig durchströmter Kugel- und

TE

D

Zylinderschüttungen, Chemie Ing. Tech. - CIT. 44 (1972) 1068–1071. doi:10.1002/cite.330441806.

D. Nemec, J. Levec, Flow through packed bed reactors: 1. Single-phase flow, Chem.

CE P

[35]

Eng. Sci. 60 (2005) 6947–6957. doi:10.1016/j.ces.2005.05.068. K.G. Allen, T.W. von Backström, D.G. Kröger, Packed bed pressure drop dependence

AC

[36]

on particle shape, size distribution, packing arrangement and roughness, Powder Technol. 246 (2013) 590–600. doi:10.1016/j.powtec.2013.06.022. [37]

E. Izadi, A. Bezuijen, Simulation of granular soil behaviour using the Bullet physics library, Geomech. from Micro to Macro. (2013) 1565–1570. http://www.crcpress.com/product/isbn/9781138027077.

[38]

M. Icardi, G. Boccardo, D.L. Marchisio, T. Tosco, R. Sethi, Pore-scale simulation of fluid flow and solute dispersion in three-dimensional porous media, Phys. Rev. E. 90 (2014) 013032. doi:10.1103/PhysRevE.90.013032.

[39]

J. van Gumster, Blender for dummies, 3rd ed., John Wiley & Sons Inc., New Jersey,

ACCEPTED MANUSCRIPT 2015. [40]

G. Boccardo, F. Augier, Y. Haroun, D. Ferr??, D.L. Marchisio, Validation of a novel

T

open-source work-flow for the simulation of packed-bed reactors, Chem. Eng. J. 279

T. Horneber, C. Rauh, A. Delgado, Fluid dynamic characterisation of porous solids in

SC R

[41]

IP

(2015) 809–820. doi:10.1016/j.cej.2015.05.032.

catalytic fixed-bed reactors, Microporous Mesoporous Mater. 154 (2012) 170–174. doi:10.1016/j.micromeso.2011.12.047.

A. Della Torre, G. Montenegro, A. Onorati, G. Tabor, CFD Characterization of

NU

[42]

MA

Pressure Drop and Heat Transfer Inside Porous Substrates, Energy Procedia. 81 (2015) 836–845. doi:10.1016/j.egypro.2015.12.093. [43]

G. Gentile, F. Manelli, an Efficient Computational Framework for the Advanced

[44]

TE

D

Modeling and Design, POLITECNICO DI MILANO, 2013. F. Lucci, A. Della Torre, J. von Rickenbach, G. Montenegro, D. Poulikakos, P.

CE P

Dimopoulos Eggenschwiler, Performance of randomized Kelvin cell structures as catalytic substrates: Mass-transfer based analysis, Chem. Eng. Sci. 112 (2014) 143–

[45]

AC

151. doi:10.1016/j.ces.2014.03.023. A.G. Dixon, M. Nijemeisland, E.H. Stitt, Systematic mesh development for 3D CFD simulation of fixed beds: Contact points study, Comput. Chem. Eng. 48 (2013) 135– 153. doi:10.1016/j.compchemeng.2012.08.011. [46]

T. Atmakidis, E.Y. Kenig, CFD-based analysis of the wall effect on the pressure drop in packed beds with moderate tube/particle diameter ratios in the laminar flow regime, Chem. Eng. J. 155 (2009) 404–410. doi:10.1016/j.cej.2009.07.057.

[47]

R. Di Felice, L.G. Gibilaro, Wall effects for the pressure drop in fixed beds, Chem. Eng. Sci. 59 (2004) 3037–3040. doi:10.1016/j.ces.2004.03.030.

[48]

N.-S. Cheng, Wall effect on pressure drop in packed beds, Powder Technol. 210

ACCEPTED MANUSCRIPT (2011) 261–266. doi:10.1016/j.powtec.2011.03.026. [49]

D. Mehta, M.C. Hawley, Wall Effect in Packed Columns, Ind. Eng. Chem. Process

A. De Klerk, Voidage variation in packed beds at small column to particle diameter

IP

[50]

T

Des. Dev. 8 (1969) 280–282. doi:10.1021/i260030a021.

[51]

SC R

ratio, AIChE J. 49 (2003) 2022–2029. doi:10.1002/aic.690490812. R. Caulkin, X. Jia, M. Fairweather, R.A. Williams, Predictions of porosity and fluid distribution through nonspherical-packed columns, AIChE J. 58 (2012) 1503–1512.

W. Reichelt, Zur Berechnung des Druckverlustes einphasig durchströmter Kugel- und

MA

[52]

NU

doi:10.1002/aic.12691.

Zylinderschüttungen, Chemie Ing. Tech. 44 (1972) 1068–1071. doi:10.1002/cite.330441806.

D

S. Hofmann, A. Bufe, G. Brenner, T. Turek, Pressure drop study on packings of

TE

[53]

differently shaped particles in milli-structured channels, Chem. Eng. Sci. 155 (2016)

[54]

CE P

376–385. doi:10.1016/j.ces.2016.08.011. V. Dolejs̆, I. Machac̆, Pressure drop during the flow of a Newtonian fluid through a

AC

fixed bed of particles, Chem. Eng. Process. Process Intensif. 34 (1995) 1–8. doi:10.1016/0255-2701(94)00566-4. [55]

B. Eisfeld, K. Schnitzlein, The influence of confining walls on the pressure drop in packed beds, Chem. Eng. Sci. 56 (2001) 4321–4329. doi:10.1016/S00092509(00)00533-9.

[56]

M.H. Pahl, Über die Kennzeichnung diskret disperser Systeme und die systematische Variation der Einflußgrößen zur Ermittlung eines allgemeingültigeren Widerstandsgesetzes der Porenströmung, University of Karlsruhe, 1975.

[57]

J. Tobiś, Influence of bed geometry on its frictional resistance under turbulent flow conditions, Chem. Eng. Sci. 55 (2000) 5359–5366. doi:10.1016/S0009-2509(00)00155-

ACCEPTED MANUSCRIPT X. M. Mayerhofer, J. Govaerts, N. Parmentier, H. Jeanmart, L. Helsen, Experimental

T

investigation of pressure drop in packed beds of irregular shaped wood particles,

CE P

TE

D

MA

NU

SC R

IP

Powder Technol. 205 (2011) 30–35. doi:10.1016/j.powtec.2010.08.006.

AC

[58]

ACCEPTED MANUSCRIPT Captions Figures

T

Figure 1: Scheme of the packing procedure.

IP

Figure 2: Scheme of the computational setup.

SC R

Figure 3: Cross section local velocity for PB reactor filled with spherical particle of dp=7.5 mm at wf =0.01 m/s.

Figure 4: Pressure drop comparison between CFD and EC for spherical particles.

NU

Figure 5: Streamlines of PB reactor filled with spheres at wf = 0.01 m/s for λ = 2 (left) and λ = 10 (right) geometries (velocity legend refers to both cases).

MA

Figure 6: Radial porosities of PB reactor filled with spheres for various λ. Figure 7: Modified friction factor predicted by CFD and Ergun correlation.

TE

D

Figure 8: Modified friction factor predicted by CFD and Nemec&Levec correlation. Figure 9: Pressure drop predicted by CFD and Allen et al. [36] correlation.

CE P

Figure 10: MRD variation in dependency of ζ at different λ and particle shape. Error bars represent min and max deviation from mean MRD at specific λ.

AC

Figure 11: CFD results for quadrilobes at wf = 0.01 m/s, λ = 10 and ζ = 4: (top-left) streamlines coloured by velocity field, (top-right) streamlines coloured by pressure field, (bottom-left) cross section of the PB reactor (upper-half => pressure filed, bottom-half => velocity field) and (bottom- right) computational mesh. Figure 12: Mean particle orientation for quadrilobe for various ζ and λ with corresponding MRD. Additionally, average tortuosities of arbitrary particles are labelled for various ζ and λ. Figure 13: Modified friction factors vs. λ at ζ = 4. Figure 14: Modified friction factors vs. λ at ζ = 1. Tables Table 1: Correlation equations.

ACCEPTED MANUSCRIPT

AC

CE P

TE

D

MA

NU

SC R

IP

T

Table 2: Fluid properties.

AC

Figure 1

CE P

TE

D

MA

NU

SC R

IP

T

ACCEPTED MANUSCRIPT

MA

NU

SC R

IP

T

ACCEPTED MANUSCRIPT

AC

CE P

TE

D

Figure 2

TE

D

MA

NU

SC R

IP

T

ACCEPTED MANUSCRIPT

AC

CE P

Figure 3

TE

D

MA

NU

SC R

IP

T

ACCEPTED MANUSCRIPT

AC

CE P

Figure 4

AC

Figure 5

CE P

TE

D

MA

NU

SC R

IP

T

ACCEPTED MANUSCRIPT

D

MA

NU

SC R

IP

T

ACCEPTED MANUSCRIPT

AC

CE P

TE

Figure 6

TE

D

MA

NU

SC R

IP

T

ACCEPTED MANUSCRIPT

AC

CE P

Figure 7

TE

D

MA

NU

SC R

IP

T

ACCEPTED MANUSCRIPT

AC

CE P

Figure 8

CE P AC

Figure 9

TE

D

MA

NU

SC R

IP

T

ACCEPTED MANUSCRIPT

AC

Figure 10

CE P

TE

D

MA

NU

SC R

IP

T

ACCEPTED MANUSCRIPT

TE

D

MA

NU

SC R

IP

T

ACCEPTED MANUSCRIPT

AC

CE P

Figure 11

CE P AC

Figure 12

TE

D

MA

NU

SC R

IP

T

ACCEPTED MANUSCRIPT

D

MA

NU

SC R

IP

T

ACCEPTED MANUSCRIPT

AC

CE P

TE

Figure 13

D

MA

NU

SC R

IP

T

ACCEPTED MANUSCRIPT

AC

CE P

TE

Figure 14

ACCEPTED MANUSCRIPT Table 1 Equation for

Wall

Modified friction factor and Reynolds

Constants

Δp / L

correction

sphere 2 (1   ) 2   w f 1    wf 150    1.75  3  f 3 2  dp  dp

sphere 2 (1   )2   w f 1    wf M   M2  3  f  3 2  dp  dp BW

M 1 

4dp

/

; BW  a1 (dp / d )2  a2 

2

6D(1   )

cylinder 2 150 (1   ) 2   w f 1.75 1    f  w f   2  4  3  3 3 3 2  d SV  d SV  

polylobed extrudates

D TE

sphere&cylinder

 

CE P

Ap  1     a1 w 2f a2     Re  Re c    f  2  4  V    3  Duct  p  Duct

AC

a2=1.5

YES

a2 = 0.88

f N&L  f * 

4

3

 ; Re N&L  Re* 

1

6

polylobed extrudates

AErgun = 150 4

fN&L  f * 2 ; Re N&L  Re* 

2 150 (1   ) 2   w f 1.75 1    f  w f   2  2  3  6 3 5  d SV   d SV 

Allen et al. [36]

NU

[35]

NO BErgun = 1.75

cylinder

MA

Nemec &Levec

AErgun = 150

c = 150

SC R

Reichelt [34]

c

T

wd Re *  f f p  (1   )

IP

Ergun [1]

(p / L) 3dp A  Ergun  BErgun (1   )  f w 2f Re*

f *

f N &L 

AErgun Re N & L

 36  Vp2  B Ergun ;    3  Ap

NO 5

1

3 6  Vp  ; d SV   Ap 

sphere

2  f * Re cDuct  Allen f Allen   a1 Re Allen  a2 ; dSV  Re*   Re Allen     dSV  Allen  Re Duct 

BErgun = 1.75

c 1

4  Vp fw   Allen;  Allen   (1   )  Sp

a1 = 172 a2=4.36 c=0.12

NO cylinder a1 = 216 a2=8.8 c=0.12

ACCEPTED MANUSCRIPT Table 2 Abbreviation

Value

T

293.15 K

Viscosity (water)

μ

0.001

Pa∙s

Density (water)

ρf

1000

kg/m3

AC

CE P

TE

D

MA

NU

SC R

IP

Temperature

T

Properties

SC R

IP

T

ACCEPTED MANUSCRIPT

AC

CE P

TE

D

MA

NU

Graphical abstract

ACCEPTED MANUSCRIPT Highlights

CE P

TE

D

MA

NU

SC R

IP

T

CFD simulations for 52 different packed bed reactor geometry set-ups were conducted A comprehensive comparison between CFD and empirical correlations (EC) was made EC give solid prediction of pressure drops for column to particle diameter > than 4 Wall influence on the pressure drop at low  (<4) is depended of particle shape Most relevant parameter for pressure drop prediction is wetted surface (laminar flow)

AC

• • • • •