Accepted Manuscript Intraparticle and interstitial flow in wide-pore superficially porous and fully porous particles Mark R. Schure, Robert S. Maier, Taylor J. Shields, Clare M. Wunder, Brian M. Wagner PII: DOI: Reference:
S0009-2509(17)30537-7 http://dx.doi.org/10.1016/j.ces.2017.08.024 CES 13767
To appear in:
Chemical Engineering Science
Received Date: Revised Date: Accepted Date:
8 June 2017 9 August 2017 27 August 2017
Please cite this article as: M.R. Schure, R.S. Maier, T.J. Shields, C.M. Wunder, B.M. Wagner, Intraparticle and interstitial flow in wide-pore superficially porous and fully porous particles, Chemical Engineering Science (2017), doi: http://dx.doi.org/10.1016/j.ces.2017.08.024
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.
Intraparticle and interstitial flow in wide-pore superficially porous and fully porous particles Mark R. Schurea,∗, Robert S. Maierb,∗, Taylor J. Shieldsc , Clare M. Wunderd , Brian M. Wagnerc a
Theoretical Separation Science Laboratory Kroungold Analytical Inc., 1299 Butler Pike, Blue Bell, PA 19422 USA b
Department of Chemical Engineering and Materials Science University of Minnesota, Minneapolis, MN 55455-0132 USA c
Advanced Materials Technology, Inc. Suite 1-K, Quillen Building, 3521 Silverside Rd, Wilmington, DE 19810 USA d
Department of Chemical and Biomolecular Engineering University of Delaware, Newark, DE 19716 USA
Abstract Using a synthetic particle model, viscous fluid flow through packed beds of porous particles is directly simulated at the pore scale (1000˚ A) using the lattice Boltzmann method to characterize intraparticle and interstitial flow. Synthetic particle models are derived from synthesis conditions, scanning electron and focused ion-beam microscopy. A fully porous particle (FPP) and a superficially porous particle (SPP), derived from the FPP, both with the same external surface, are studied. Packed beds of random packings and body-centered cubic packings of the SPP and FPP models were generated by a Monte Carlo procedure that employs random translation and rigid-body rotation of the particles. Detailed velocity distributions are presented for the interstitial and intraparticle regions of the packed beds and within the particles. These results confirm that porous particle packed beds are heterogeneous systems which require extensions to classical theory for correctly predicting the resistance to flow. It is shown that SPPs require less pressure than FPPs to maintain the same flow velocity. For the random SPP and FPP packed beds, the particle hull mass flux is ≈ 10% of the interstitial flux and ≈ 3% of the total volumetric flux in the flow direction through the SPP hull. The calculated intraparticle pore velocities confirm that an internal flow, characteristic of “perfusion” chromatography, exists within the porous shell that can enhance biomolecular separations.
∗
Corresponding authors, e-mail:
[email protected] and
[email protected]
Preprint submitted to Elsevier
August 29, 2017
1
1. Introduction
2
Many of the newer particles utilized in liquid chromatograhy (LC) use core-shell particles, also
3
known as superficially porous particles (SPPs) (DeStefano et al., 2008; Schuster et al., 2012; Wagner
4
et al., 2012; Schuster et al., 2013; Schure and Moran, 2017; Wagner et al., 2017). These particles
5
contain a nonporous core with a porous shell on the outside. This is in contrast to fully porous
6
particles (FPPs) which have been traditionally used in LC, and “pellicular” particles (Horv´ath and
7
Lipsky, 1969), which have thicker shells than modern SPPs and were introduced in the late 1960s.
8
Many questions still exist about SPPs, including understanding the flow-field inside and outside of
9
the particle in wide-pore SPP materials. The work reported herein attempts to explain some of these
10
questions.
11
A number of large-scale simulation studies have examined the microscopic details of solute trans-
12
port through a packed bed of nonporous particles (Maier et al., 1998, 2000; Kandhai et al., 2002;
13
Maier et al., 2003; Bijeljic et al., 2004; Sullivan et al., 2005). This technology has also been used
14
to study chromatographic column processes (Schure et al., 2002; Schure and Maier, 2006; Hlushkou
15
et al., 2007; Khirevich, 2011). In all of these studies, the chromatographic particle is a nonporous
16
sphere enabling the study of fluid-phase mass transport; the positions of these particles are explicitly
17
stated in the packed bed by using particle packing software to construct the bed.
18
No transport models have been developed previously that explicitly define a detailed pore structure
19
at the particle level. In cases where transport is to be discussed for porous particles, stochastic, coarse-
20
grained effects are utilized to predict transport under random diffusion (Spaid and Phelan Jr, 1997;
21
Kandhai et al., 2002; Daneyko et al., 2015). In this and all other scenarios used in simulating packed
22
beds for studying LC, it has been assumed that fluid flow ceases in the particle phase as the pore size
23
is much too small, typically ≤ 400˚ A, to permit flow through any of the particle pore structure. This
24
would justify the assumption of a purely diffusive transport of solutes into and out of the intraparticle
25
pore space.
26
It is well-recognized that particles with large connecting porous networks, often with pore sizes
27
on the order of ≥ 5000˚ A, have a significant internal convective flow (Lloyd and Warner, 1990; Afeyan
28
et al., 1991; Rodrigues et al., 1993; Frey et al., 1993; Rodrigues, 1997; de Neuville et al., 2014; Wu
29
et al., 2015); this is often referred to as “perfusion” and is characteristic of wide-pore materials. In one
30
of these works (Frey et al., 1993) particles were referred to as “gigaporous” when the pore diameter
31
to particle diameter ratio is > 0.01. The particles studied here fit that definition. 2
32
Previous studies have not examined the microfluidic details of the surface and internal flow of wide-
33
pore particles with pore diameters ≥ 1000˚ A. If there is flow into (and out of) wide-pore particles, this
34
would enhance diffusion. The extent of convective flow in and around the pore surface of a porous
35
particle is extremely hard to ascertain by experiment, although single particle measurements have
36
been made (Pfeiffer et al., 1996).
37
A number of investigations have examined the internal flow in porous particles. For example,
38
in a well-known study (Neale et al., 1973), Brinkman’s equation (Brinkman, 1947, 1949; Durlofsky
39
and Brady, 1987) was used as the basis to predict the permeability of porous particle beds. These
40
authors calculated bed permeability as a function of the interstitial porosity for various values of β, the
41
nondimensional particle permeability. Their results show that for β > 100, the particle permeability
42
has little effect on the bed permeability, and this is the range in which the simulations in this paper
43
are relevant.
44
Guiochon and coworkers (Gritti et al., 2007) measured the porosity and permeability of SPP
45
and FPP packed columns. Although they found the SPP column less permeable than the FPP, the
46
differences in diameter and interstitial porosity of their SPP and FPP columns complicate a direct
47
comparison. The Kozeny-Carman (KC) equation (Carman, 1937; Giddings, 1965; Neue, 1997; Quinn,
48
2014) was also used to predict permeability based on particle size and interstitial porosity. (Note
49
that this approach would predict the same permeability for SPP and FPP particles of the same
50
diameter in columns of the same interstitial porosity.) They found the SPP column permeability less
51
than predicted by theory and less than the prediction for the FPP column. They suggested that
52
the greater rugosity of the SPP, which is a measure of the deviation from a purely spherical surface
53
and also referred to as surface roughness, might play a role in understanding the permeability, but
54
they did not propose a specific physical explanation. Yet in another study (Ismail et al., 2016), the
55
SPP and FPP particle diameters and interstitial porosities were more similar than those studied by
56
Guiochon and coworkers, mentioned above, and it was found that the SPP column was 22% more
57
permeable than the FPP column.
58
A porous particle can be regarded as a rough sphere by ignoring intraparticle flow. The effect
59
of surface roughness on packed bed pressure drop has been studied in several contexts. It has been
60
found (Crawford and Plumb, 1986) that the pressure drop increases with increasing surface area
61
and roughness of nonporous particles. However, it was demonstrated (Eisfeld and Schnitzlein, 2001)
62
that increased porosity associated with rough particles counteracted this effect. Others (Nemec and 3
63
Levec, 2005; Allen et al., 2013) have suggested that this effect can be attributed to departures from
64
sphericity rather than surface roughness. Nonetheless, the surface roughness literature highlights the
65
importance of accounting for porosity and surface area within the porous particle structure.
66
The influence of SPP morphology on pore-level flow can be studied with high resolution fluid
67
mechanics if a model of the SPP or FPP can be formulated. For studying wide-pore materials,
68
computer simulation of the model can yield useful and realistic information on fluid flow. One of
69
the benefits of simulation is that differences can be eliminated in particle surface geometry when
70
comparing SPP and FPP packed beds. This makes it possible to evaluate theoretical approaches,
71
such as the KC equation, for predicting porous particle flow properties. Differences in permeability
72
between SPP, FPP and solid-sphere beds are discussed in the work presented here, including the
73
contrast of some theoretical predictions with calculations reported herein.
74
In this paper a synthetic particle model derived from synthesis conditions, electron microscopy
75
and particle simulation techniques, described below, are used to study the detailed flow near, into and
76
out of the particle. Both SPP and FPP morphologies can be produced as sphere models, which are
77
included as supplemental files. The surface area and pore volume of the sphere models are in close
78
agreement with experimental particles made in the laboratory. Calculation of a high-resolution flow
79
field in the interstitial and intraparticle pore regions of the models is therefore considered a simulation
80
of flow through the experimental particles. Specific attention is given here to understanding the
81
permeability, mass and volumetric flux and detailed flow field that exist both internal and external
82
to the particle.
83
2. Experimental and Computational Methods for Particles
84
2.1. Experimental methods and techniques
85
A number of silica SPPs were prepared using the layer-by-layer method (Hayes et al., 2014) to put
86
porous shells on nonporous cores. The porous shell is composed of silica sol particles with average
87
radius rsol = 0.0575 µm. The average core radius rc = 1.65 µm and the thickness of the shell, Ls , is
88
≈ 0.58 µm. These parameters are utilized for the construction of the synthetic particle, as described
89
below in section 2.2. The particle diameter is the mean value of 70 particles, each with two orthogonal
90
measurements taken utilizing scanning electron microscopy (SEM) (Reimer, 2013). SEM images were
91
obtained using a Zeiss (Jena, Germany) Auriga 60 high resolution SEM with integrated focused ion
92
beam (FIB) (Yao, 2007) at the University of Delaware (Newark, DE).
4
93
A number of these parameters are summarized in Table 1 for both experimental and synthetic
94
particles. For the experimental particles, dividing the standard deviation by the mean gives a relative
95
variation in mean diameter of 4.6%, a relatively narrow size distribution. One of the SPP systems
96
was chosen for further study, with label 1030-27-D-SPP, a 1000˚ A mean pore diameter SPP.
97
Experimental surface areas and pore volumes were measured with a Micromeritics (Norcross, GA)
98
Tristar II instrument using the N2 BET adsorption method for surface area determination (Brunauer
99
et al., 1938) and the BJH method for pore volume determination (Barrett et al., 1951). The pore
100
volume was determined using the desorption branch of the N2 isotherm. The method of calculating
101
the surface area, porosity and hull radius of synthetic particles is described below in section 2.3. The
102
synthetic particle hull radius is rh ≈ 2.15 µm for all particles in Table 1. The agreement between
103
synthetic and experimental particle diameter is within 3.7% which was deemed acceptable.
104
Two experimental views of the SPP are shown in Figure 1 using SEM and FIB for the cross-
105
sectional view. Porosity calculations from FIB slice image analysis data predict an average porosity
106
of ≈ 50%. This was the starting point for synthetic particle construction.
107
2.2. Synthetic particle construction
108
Synthetic SPP and FPP particles are created by a two-step process. First, a homogeneous random
109
packing of sol particles is generated, sufficiently large to contain the SPP or FPP hull. Second, the
110
SPP core particle is inserted into this bulk phase, and sol particles lying outside the SPP or FPP hull
111
or inside the core particle are removed.
112
The bulk phase can be created by any method for generating random, close-packed spheres. In
113
the present work, 27,000 sol particles were gradually compressed within periodic boundary conditions
114
(PBCs) to create cube-shaped packs of specified porosity (50%, 45% and 36%). Since these packs
115
were slightly smaller than the SPP hull, periodic images of the packs were tiled in a periodic array
116
to create a bulk phase, or superstructure, that would completely contain the SPP or FPP hull. Any
117
sol particles that intersected or lay outside the hull were then removed, if the radial distance from
118
the center satisfied r > rc + LS − rsol . For the SPP, a core particle was inserted at the center of the
119
superstructure, and sol particles were removed where r < rc . Sol particles satisfying rc < r < rc + rsol
120
were retained in the model, giving the core particle a rough texture. The synthetic sol particle radius
121
for all SPP and FPP is rsol = 0.0575 µm and the synthetic core radius of the SPP is rc = 1.65 µm
122
corresponding to experimental values.
123
Defects in the particle surface are then introduced by selectively removing particles from the outer 5
Experimental particle designation
Surface area m2 /g
Pore volume cc/g
Layers
Diameter µm
Stnd. Dev. µm
1030-27-D-SPP
7.94
0.230
≈4
4.46
0.206
Synthetic particle designation
Surface area m2 /g Sm
Pore volume cc/g Vp,m
Surface area % deviation from experiment
Pore volume % deviation from experiment
Intraparticle porosity sp
1030-27-D-50-0-SPP 1030-27-D-50-0-FPP
8.26 22.4
0.200 0.470
-3.97
11.9
0.560 0.510
1030-27-D-45-0-SPP 1030-27-D-45-0-FPP
8.75 22.5
0.180 0.390
-10.2
20.6
0.520 0.460
1030-27-D-36-0-SPP 1030-27-D-36-0-FPP
9.47 22.6
0.150 0.280
-19.3
35.7
0.450 0.382
1030-27-D-50-10-SPP 1030-27-D-50-10-FPP
7.74 21.8
0.220 0.492
2.56
3.76
0.601 0.521
1030-27-D-45-10-SPP 1030-27-D-45-10-FPP
8.22 21.9
0.200 0.410
-3.56
12.2
0.560 0.470
1030-27-D-36-10-SPP 1030-27-D-36-10-FPP
8.92 21.9
0.170 0.290
-12.4
26.7
0.490 0.390
Table 1: Properties of the experimental and synthetic models for this study. The first three numbers and letters (103027-D) in the particle designation are the lab notebook entry of the experimental particle. The next two entries for the synthetic particle are the 27,000 particle cubic pack interstitial porosity (as a percentage) used in particle building and the percentage of particles removed from the first two outer layers. The abbreviation SPP and FPP are as per the text. The porosities, sp , are determined using the random sampling algorithm described in the text.
6
124
surface inwards. A number of removal algorithms have been developed that vary in sophistication.
125
Both the SPP and FPP structures are produced simultaneously so that both contain exactly the same
126
surface defects. All software is written in FORTRAN-90, C++ and MATLAB for calculations and the
127
POVRay scripting language is utilized for particle visualization.
128 129
2.3. Synthetic particle metrics 2.3.1. Hull radius
130
The hull radius is an artificial boundary used to separate the interstitial region from the particle
131
region. The hull radius rh is estimated by sorting the distance of the sol sphere centers from the
132
origin and storing the 300 largest distances. To these distances, rsol is added, and the mean and
133
standard deviation of these are determined. The mean of these distances is the hull radius, rh , which
134
is ≈ 2.15 µm. The standard deviation is less than one-thousandth of the mean.
135
2.3.2. Porosity
136
The intraparticle porosity is the ratio of porous volume to a specified particle volume. When
137
discussing FPPs, the intraparticle porous volume is unambiguous. With SPPs there are two ratios:
138
one which uses the hull volume, Vh = 43 πrh3 , and the other includes just the shell volume, Vs = Vh − Vc ,
139
where the core volume is Vc = 43 πrc3 . The latter is necessary to relate the experimental surface area
140
and pore volume to that of the model particles. We use the symbol p when the specified volume is
141
the hull volume and sp when the shell volume is used as the specified volume. Note that for FPPs,
142
the shell volume is the hull volume and sp = p . Another intraparticle porosity, 0p , used in section
143
2.5, is the ratio of the total pore volume in all particles of the packed bed to the total box (column)
144
volume.
145
146
The intraparticle porosity, p , may be computed as the complement of the ratio of solid volume of the SPP or FPP, identified as Vt,s , to hull volume Vh as:
p = 1 −
Vt,s Vh
(1)
147
3 3 For FPPs Vt,s = nsol 43 πrsol and for SPPs Vt,s = nsol 43 πrsol + Vc . This works fine for sol particles with
148
no overlap, as is the case with FPPs.
149
Alternatively, p can be determined for FPPs and SPPs by choosing random points in the particle
150
volume and taking the fraction of points not within the solid material. Typically 500,000 random
151
points are chosen with the hull volume. For SPPs, sp is determined by sampling just the shell region
7
152
to determine the pore volume. Typically, 100,000 random points are sampled between the hull and
153
core region for sp .
154
2.3.3. Pore volume
155
The mass specific pore volume, Vp,m , is:
Vp,m
sp · Vs = (Vc · ρc ) + (1 − sp ) · Vs · ρs
(2)
156
where ρc and ρs are the density of the core and shell, respectively. For FPPs Vc = 0, Vs = Vh and
157
sp = p . If the density of the core and shell are equal, ρc = ρs = ρcs , and Equation 2 reduces to
Vp,m =
sp · Vs (Vc + (1 − sp ) · Vs ) · ρcs
158
where the density of the solid material ρcs = 2.20 g/cc (Iler, 1979).
159
2.3.4. Surface area
(3)
160
The total surface area of the particle structure is the sum of all of the individual sphere surface
161
areas corrected for any overlap. If a chosen sphere has overlapping spheres in its volume, only the
162
non-overlapped area is included in the calculation of surface area. The procedure for determining the
163
overlapped area includes tiling 10242 nearly equispaced points on each sphere surface (von Laven,
164
2015) and determining the fraction of points, ξi , which are not included in any overlapping sphere
165
volume. The mass specific surface area is n P
Sm =
Si · ξi
i=1
(Vc + (1 − sp ) · Vs ) · ρcs
(4)
166
Again, for FPPs Vc = 0, Vs = Vh and sp = p . The surface area per particle is given as Si = 4πri2
167
noting that the porous particle is comprised of n spheres including the core sphere when present.
168
2.4. Model particles
169
The comparison of the experimental and synthetic particles given in Table 1 shows that the
170
synthetic particle 1030-27-D-50-10-SPP closely matches the experimental particle surface area and
171
porosity. This supports investigation into the detailed microfluidic environment around and in these
172
particles, which is described below. The visual agreement in Figure 1 between experimental and
173
simulated particles was further justification to use these models for fluid mechanics calculations.
174
Although particle models cannot be produced that match exactly with the experimental particle, at 8
175
the sol particle level, this ultimate accuracy may not be required to produce a model sufficiently
176
realistic enough to study flow in wide-pore porous materials.
177
The SPP has 12,209 sol particles plus one core particle, and the corresponding FPP particle
178
contains 24,014 randomly packed sol particles. The sol particles in the outer layers of the FPP have
179
the same etching step as the SPP and are arranged identically to the SPP. The internal porosity
180
of the SPP particle is 31.2% when the core is included in the particle volume. The porosity of the
181
FPP particle is 53.9% using Equation 1. This corresponds closely to the 50% native pack porosity
182
that was used to construct the SPP. Removing sol particles from the surface results in increasing the
183
particle porosity. There is typically ≈ 3% or less difference between sp calculated using Equation 1
184
and the randomly sampled particle algorithm described above; the number of sampling points appears
185
to make only minor differences.
186
2.4.1. Local particle porosity profile
187
Figure 2 shows the local particle porosity in the SPP and FPP as a function of radial distance
188
from the particle center. The two particles have the same sol geometry in the region surrounding
189
the SPP core. Hence, the local particle porosity profiles are identical just beyond the core radius.
190
The SPP local porosity is elevated at the core because some sol particles that intersect the core were
191
removed, leaving slightly larger pore spaces. The FPP local porosity fluctuates near the center due
192
to the small volume of the radial sampling regions.
193
The inset of Figure 2 shows the local porosity profile in the region close to the hull, located at
194
r ≈ rh . The local porosity decreases from unity at the hull to a value of ≈ 0.5 at a radial position
195
one sol diameter from the hull. If the effective radius of the SPP were decreased by one sol particle
196
radius, i.e. reff = rh − rsol , the volume excluded by reducing the radius would be roughly 90% pore
197
and 10% solid. This volume would be the equivalent of 350 sol particles or about 3% of the total sol
198
mass. In this case, the volume would be considered part of the interstitial region, and the interstitial
199
porosity, , would increase from 34.1% (based on rh ) to 38.8% (based on reff ). Figure 2 can be used
200
to estimate how the intraparticle and interstitial local porosities would change if the hull radius were
201
defined differently.
202
2.5. Construction of synthetic particle packed beds
203
Packed beds of porous particles are constructed by a two-step process. First, a packed bed of
204
spheres is generated by a hard-sphere Monte Carlo method (Tildesley and Allen, 1987). Second, each
9
205
sphere in the packed bed is replaced by a porous particle of equivalent hull radius, and the bed is
206
then further compressed by a similar Monte Carlo method.
207
The first step in the process has been described in earlier work (Maier et al., 2003). Briefly, each
208
major iteration performs one or more Monte Carlo sweeps in which each sphere is assigned a random
209
translation. The trial move is accepted if it does not result in overlap. The sweeps are followed by
210
compression (scaling) of the bounding box, which concludes the major iteration. The bounding box
211
may have a combination of impenetrable and periodic faces, or it can be described by some other
212
simple geometry, such as a cylinder.
213
The second step of the process begins by replacing the packed spheres with SPPs or FPPs. Each
214
porous particle is given a random rigid-body rotation that preserves the internal geometry of the
215
sol particles. Because the porous particle hull is a bounding sphere around the sol particles, simply
216
substituting them for solid spheres leaves loose contacts between neighboring particles. Therefore,
217
further iterations of the Monte Carlo method are performed on the porous particle packed bed.
218
The random move for each particle consists of both a rigid-body rotation and translation. In-
219
dividual pairs of sol particles belonging to neighboring porous particles must be tested for overlap.
220
Overlap testing is made more efficient by restricting attention to sol particles that populate the exte-
221
rior of the particles. During these iterations, the porous particles undergo relatively small rotations
222
and translations, and the compression typically eliminates gaps on the order of a sol radius between
223
neighboring particles. Although the neighboring particle hulls overlap slightly, the sol particles do
224
not. Because the relative particle motion is so limited, the interstitial region of the resulting porous
225
particle bed remains similar to the original packed bed of spheres.
226
A random packing of porous particles was simulated by replacing the spheres in a close-packed
227
bed of 125 solid spheres contained in a cubic bounding box with periodic boundaries. This is shown
228
in Figure 3A. The dimension of the original bounding box is 9.35 · rh with 36% porosity. After further
229
Monte Carlo compression, the bounding box dimension was reduced by 1% or ≈ 3 sol radii. The
230
resulting interstitial region is 34% of the bounding box volume, compared with 36% shown in Figure
231
3B.
233
A body-centered cubic (BCC) packing of porous particles was constructed by placing a porous √ particle at the center of a BCC unit cell. The initial unit cell size was 4 · rh / 3. One porous particle
234
is located at the center and one at each unit cell corner. After compression, the unit cell size was
235
reduced by ≈ 1%, slightly more than one sol radius. The resulting interstitial region is 29% of the
232
10
236
unit cell volume, compared with 32% for the original unit cell.
237
A question remains whether the SPP or FPP could be packed more densely. This is difficult to
238
assess because the porous particle surface is not uniform, but it would depend to a certain extent on
239
the ability of the porous particle to rotate into more compact positions. It is possible that starting with
240
a looser bed of solid spheres would give the SPP or FPP room to explore more compact orientations.
241
2.5.1. Packed bed porosity
242
Table 2 summarizes the box size, the number of particle structures present in the bed and the
243
three porosity values characteristic of the solid-sphere and porous particle packed beds utilized in the
244
following calculations. The total porosity, T , is the sum of the intraparticle porosity, 0p , which is
245
the ratio of void space in the particle shells to the total bed volume, and the interstitial porosity, ,
246
which is the ratio of the volume between particle hulls and the total bed volume of the column. For
247
a solid-sphere packing, the bed porosity and interstitial porosity are the same.
248
Porous particles substantially increase the total porosity. The total porosity of the random bed
249
solid-sphere packing is 36%, which is the random close packing limit (Torquato et al., 2000; Song
250
et al., 2008) for spheres. By replacing the solid spheres with SPPs, the total porosity increases to
251
55%. By replacing the solid spheres with FPPs, the total porosity increases to 70%. These increases
252
in total porosity are due primarily to the void space in the porous particles, since their interstitial
253
volume differs by only 2% to 3% from the solid-sphere bed.
254
Although porous particles increase the total bed porosity, their void volume does not dominate the
255
interstitial volume. For the random SPP packed bed, the intraparticle and interstitial pore volumes
256
are 38% and 62% of the total bed volume, respectively. For the random FPP packed bed, the
257
corresponding ratios are 49% and 51%. Increasing the sol packing density within the porous particle
258
is not likely to have much impact on these ratios. Assuming the sol could be packed to 64% density,
259
the intraparticle share of the total bed volume would increase by 3% and 4% for the SPP and FPP,
260
respectively.
261
2.5.2. Nearest-surface distributions
262
The nearest-surface distribution (NSD) is a probability density function for the distance from
263
a point in the fluid phase to the nearest solid surface and is an indicator of the pore size. The
264
NSD for the random SPP packed bed is shown in Figure 4. The NSD is separated into two parts;
265
an intraparticle and an interstitial distribution. The combined area beneath the intraparticle and
266
interstitial distributions is equal to unity, based on relative pore volumes of 38% and 62%, respectively. 11
Packing
Box size µm
random SPP random FPP random sphere BCC SPP BCC FPP BCC sphere
19.89 19.89 20.08 4.896 4.896 4.959
Number Total of porosity particles T 125 125 125 2 2 2
0.550 0.696 0.360 0.517 0.674 0.320
Bed porosity interstitial intraparticle 0p 0.341 0.341 0.360 0.294 0.294 0.320
0.208 0.355 0.000 0.223 0.380 0.000
Table 2: The bed packing designations used in this study derived from 1030-27-D-50-10-SPP and 1030-27-D-50-10-FPP.
267
However, the two distributions are nearly disjoint, reflecting the different characteristic length scales
268
of the particle pore structure, on the order of dsol , and the particle diameter, on the order of 2rh .
269
The intraparticle distribution ranges from zero to approximately one sol diameter. This range is
270
much wider than a homogeneous and infinite random sphere packing where the upper limit is closer
271
to a sol particle radius. The intraparticle distribution has a higher limit here because the hull forms
272
an open boundary and the distance from the hull to the nearest sol particle, in many cases, exceeds
273
a sol particle radius.
274
The intraparticle distributions for SPP and FPP are compared in the inset to Figure 4. The area
275
under the FPP curve is normalized to unity, while the area under the SPP curve reflects its lesser
276
pore volume. Both the FPP and SPP have the same large-pore distributions because the outermost
277
region of the particles, near the hull, are identical. The FPP distribution has a higher density of small
278
pores than the SPP because it has a porous region of close-packed sol corresponding to the solid SPP
279
core region.
280
The SPP interstitial distribution in Figure 4 is also representative of the FPP interstitial distri-
281
bution, since the two packed beds have identical construction. In the interstitial region, the distance
282
to the nearest sol particle ranges from zero to approximately eight sol diameters. Note the interstitial
283
distribution is actually an approximation. It is based on the NSD for the solid-sphere random pack-
284
ing, which gives the distance from interstitial points in the fluid to the particle hull surfaces. The
285
additional distance from the hull surfaces to the nearest sol particles was approximated by shifting
286
this NSD one half sol diameter to the right (roughly the average distance from the hull to the nearest
287
sol particle). This shift is responsible for the discontinuity in the distribution shown in Figure 4. The
288
true interstitial distribution would be continuous and increasing from zero to a peak near half of a 12
289
sol diameter, while the remainder of the true distribution would be similar to the approximation.
290
3. Fluid mechanics
291
3.1. Lattice Boltzmann method
292
Fluid motion in the pore space is described by the incompressible Navier-Stokes equations with
293
appropriate initial and boundary conditions (White and Corfield, 2006). Fluid flow was simulated
294
under the assumptions of steady, saturated flow at small Mach number using the lattice Boltzmann
295
(LB) method. The LB method solves the discrete-velocity Boltzmann equation for the fluid mass and
296
momentum density distributions (Sterling and Chen, 1996; Maier et al., 1998; Maier and Bernard,
297
2010; Kr¨ uger et al., 2016). The method is a pseudo-transient, explicit scheme for recovering Navier-
298
Stokes behavior in the low Mach-number limit. The LB equations are solved on a regular 3-D grid
299
with the D3Q19 method (Maier et al., 1998; Wagner, 2008; Kr¨ uger et al., 2016), in which physical
300
space is discretized on a computational grid and velocity space is represented at each point by a set
301
of 19 direction vectors. The computational grid is then superimposed on the packed bed geometry.
302
The LB relaxation parameter which controls the rate at which local equilibrium is imposed at
303
each time step was set to unity. This parameter establishes the ratio of the lattice grid spacing
304
and time step in the method, typically referred to as the “lattice viscosity.” The desired physical
305
fluid parameters, including the dynamic viscosity, 1.0 × 10−3 m2 s−1 , are obtained by setting the grid
306
spacing to achieve the necessary spatial resolution (see below) and solving for the time step length
307
that makes the lattice and physical Reynolds numbers equivalent.
308
The method is made parallel by domain decomposition of the computational grid; sub-domains
309
are mapped to computer processors using the MPI Cartesian communicator topology (MPI Forum,
310
2016; Gropp et al., 1999). Border cell data is exchanged between adjacent sub-domains at each LB
311
time step. Only grid cells within the pore space are part of the flow calculation.
312
Flow was driven by a uniform pressure gradient, and no-slip conditions were enforced at all solid
313
boundaries. Gravitational forces were not included. PBCs were used to wrap the flow around to the
314
opposite face of the simulation box. The flow simulation was initiated at zero velocity and iterated to
315
a steady state. Application of the method to porous media flows and its accuracy has been evaluated
316
in a previous work (Maier and Bernard, 2010).
317
The samples of porous media simulated in this work are relatively small because finely-spaced
318
computational grids are required to accurately model fluid dynamics in the pore spaces. This is
319
especially true for porous particles because the flow must be resolved in the sol pore space and in the 13
320
interstitial region surrounding the particles. Periodic boundaries are therefore an attractive method
321
for simulating flows in small samples without the entrance or exit effects associated with external
322
boundaries. Although the sample is periodic on the length scale of the box, there is no repetition of
323
macropore geometry within the sample, and each porous particle has a random rotational orientation.
324
The flow simulations are intended to describe velocity distributions within and around the particles.
325
Previous results show that for monodisperse spheres of diameter d, the velocity distribution becomes
326
well-converged when the grid spacing, ∆x, is decreased from d/20 to d/40 as shown in Figure 13, p.248 of
327
(Maier and Bernard, 2010). In the present work, ∆x = dsol /24, where dsol is the sol particle diameter.
328
Thus, the velocity distributions for the sol pore space have adequate resolution, while distributions
329
for the larger macropore spaces have higher resolution. Flow was simulated in the random SPP and
330
FPP packs using grid dimensions of 4150 × 4150 × 4150. The BCC SPP and FPP packs use grid
331
dimensions of 1024 × 1024 × 1024, the random sphere pack uses grid dimensions of 180 × 180 × 180
332
and the BCC sphere pack has dimensions of 176 × 176 × 176. These packs are described in Table 2.
333
3.2. Pressure and permeability
334
The particle Reynolds number is defined as Resol = vdsol /ν, where v is the mean pore velocity and
335
ν the kinematic viscosity. Flows were simulated at Resol = 1 × 10−3 . The Darcy flux, or superficial
336
velocity, is defined as q = εv. The permeability, k, as defined in Darcy’s Law (Brinkman, 1947;
337
Giddings, 1965; Bear, 2014; Quinn, 2014) is q = −(k/µ)∇p, where ∇p is the pressure gradient and µ
338
is the dynamic viscosity. In this work, the pressure gradient is always aligned with the z-axis, so the
339
permeability is calculated simply as k = qz µ/(dp/dz).
340
Higher pressure is required to drive the mean pore velocity at constant Resol with porous particles.
341
For the random packings, the required pressure is about 44% (SPP) and 100% (FPP) higher than the
342
solid-sphere packing, as shown in Table 3. For the SPP and FPP BCC packings, the pressure is 36%
343
and 96% higher than the BCC solid-sphere packing, respectively. Considering both types of porous
344
particles and both packing arrangements, SPPs require less pressure than FPPs to maintain the same
345
flow velocity conditions.
346
347
Although the required pressure is higher for porous particles, it is not nearly so high as predicted by the KC equation, given by
dp/dz = [180µ(1 − T )2 qz ]/[d2eff 3T ] 348
(5)
for spherical particles of diameter deff . While the required pressure in the solid-sphere packing is well 14
Packing
qz (m/s)
random SPP 4.75 × 10−3 random FPP 6.05 × 10−3 random sphere 3.13 × 10−3 BCC SPP 4.47 × 10−3 BCC FPP 5.86 × 10−3 BCC sphere 2.78 × 10−3
dp/dz (kg/m2 s2 )
k (m2 )
3.78 × 108 5.31 × 108 2.62 × 108 4.16 × 108 6.01 × 108 3.06 × 108
1.26 × 10−14 1.14 × 10−14 1.20 × 10−14 1.07 × 10−14 9.75 × 10−15 9.10 × 10−15
Nondimensional average pore velocity: particle interstitial 0.067 0.059 0 0.069 0.059 0
1.57 1.98 1.00 1.71 2.22 1.00
Table 3: Bed velocity, qz , pressure gradient, dp/dz, permeability from Darcy’s equation, k, and nondimensional pore velocity in the direction of the pressure gradient for simulated solid and porous particle packed beds. All cases are simulated at Resol = 1.00 × 10−3 and the average pore velocity in the direction of the pressure gradient is v z = 8.694 × 10−3 m/s. The average pore velocity within the porous particles and in the interstitial region is made nondimensional by dividing by v z .
349
predicted by Equation 5, the pressures in the SPP and FPP packings are far lower than predicted
350
(see Table 4). These predictions used deff = 6/SV , where SV is the ratio of total surface area to
351
solid volume in the bed. The departure of the simulation results from the predictions indicates
352
that packings of porous particles are not equivalent to homogeneous packings of sol particles. Porous
353
particles admit flow preferentially through the interstitial region and are therefore better characterized
354
as heterogeneous media. The homogeneity assumption underlying Equation 5 has been discussed in
355
related works (Schlueter and Witherspoon, 1994). Packing p random FPP random SPP excl. core random SPP incl. core BCC FPP BCC SPP excl. core BCC SPP incl. core
0.540 0.571 0.310 0.540 0.571 0.310
Particle deff 1.15 × 10−7 3.16 × 10−7 3.16 × 10−7 1.15 × 10−7 3.16 × 10−7 3.16 × 10−7
Bed kp
kN EN
k
5.47 × 10−17 5.63 × 10−16 3.48 × 10−17 5.47 × 10−17 5.63 × 10−16 3.48 × 10−17
1.12 × 10−14 1.35 × 10−14 1.10 × 10−14 6.36 × 10−15 8.10 × 10−15 6.24 × 10−15
1.14 × 10−14 1.26 × 10−14 1.26 × 10−14 9.75 × 10−15 1.07 × 10−14 1.07 × 10−14
Table 4: Comparison of particle and bed permeabilities. The simulation permeability, k, theoretical permeability with Brinkman’s correction (Neale et al., 1973), kN EN , and particle permeability, kp , using the KC equation are given with both random and BCC packing. The particle porosity, p , and the effective diameter of the sol particles, deff , taking into account the surface area of the core, are used to calculate kp .
356
The permeabilities of the SPP and FPP packed beds do not differ greatly from the solid-sphere
357
packed bed given in Table 3. On one hand, this is surprising because the porous particles have a
358
much higher specific surface area which correlates with increased resistance to flow in a homogeneous 15
359
particle packing. It is unsurprising, however, because the interstitial porosity of the packs are similar,
360
as shown in Table 2, and therefore can accommodate similar volumes of flow at the same pressure.
361
Even so, the similarity is limited.
362
Previously, it was noted that the gap between the porous particle hull and the sol particles is on
363
the order of the sol radius. This additional void space between the particles allows the SPP and FPP
364
to accommodate more flow through the interstitial region than their interstitial void fraction would
365
otherwise suggest. This capacity advantage is counterbalanced by the need to pass a greater volume
366
of fluid through the interstitial region to achieve the same superficial velocity as the sphere packing.
367
The additional fluid in the particles cannot be pushed through the particle quickly, and it is instead
368
pushed around the particles at a faster rate.
369
It is also interesting to note that the SPP bed permeability is slightly greater than the solid-sphere
370
bed, while the FPP bed permeability is slightly lower. The SPP bed permeability is approximately
371
10% greater than the FPP. A similar result was reported (Ismail et al., 2016) where the SPP column
372
permeability was 22% greater than the FPP. The balance between increased interstitial capacity and
373
flow volume thus works slightly in favor of the SPP packing and slightly against the FPP packing.
374
But the advantage is small and depends on the specific packing density of the SPP particles relative
375
to the solid spheres.
376
Theoretical predictions of FPP permeability using the approach in (Neale et al., 1973) are accurate
377
in the present case for the random bed. Table 4 compares the simulated and predicted bed perme-
378
abilities and these differ by < 2% for the random FPP packing. The prediction for the BCC FPP
379
packing is less accurate, differing by 25% from the simulation. Theoretical predictions of permeability
380
are known to be less accurate for BCC packings than for random packings. In a well-known work on
381
flow in crystalline packings (Zick and Homsy, 1982), the drag coefficient for BCC solid spheres differs
382
by over 25% from the KC equation, but it agrees within 2% with our result in Table 3. Note that the
383
drag coefficient for BCC spheres is given as FD = (2/9)rh2 /(1 − )/162.9.
384
The prediction of SPP random bed permeability is more problematic than the FPP case because
385
estimation of kN EN depends on an estimate of the particle permeability. For an FPP composed of
386
monodisperse sol, it is straightforward to compute kp from the sol diameter and particle porosity.
387
For a SPP, the porosity can be defined either to include or exclude the core particle, leading to
388
different values of both the particle and bed permeability. Table 4 gives predictions of the SPP
389
bed permeability for both of these cases. Excluding the core particle volume results in a 7% over16
390
prediction of the random SPP bed permeability, while including it leads to a 12% under-prediction;
391
these are compared to the simulation value k. The former prediction is preferred, and the rationale
392
for excluding the core particle when computing the internal particle porosity is that the KC equation
393
is not intended to predict the permeability of a heterogeneous system. The calculation of permeability
394
for a SPP with Brinkman-level correction has been presented, (Masliyah et al., 1987) but was not
395
computed for this work.
396
3.3. Fluid velocity distributions
397
The fluid velocity distribution in a packed bed is a density function describing the relative mass
398
(or volume) of fluid moving at a given velocity, shown in Figure 5. In a packed bed, most of the fluid
399
is in close proximity to a solid surface. Assuming a pressure-driven flow, this fluid has relatively low
400
velocity. Therefore, the velocity distribution in a packed bed typically has a sharp peak (mode) near
401
v = 0 and decreases exponentially with increasing v. Packed beds of porous particles have the same
402
general distribution as solid-sphere beds, but the mode is more pronounced because the surface area
403
is large compared to a solid-particle bed.
404
In porous particle beds, all solid surfaces are internal to the particles because the hull is defined as
405
a sphere surrounding the sol with radius rh . Therefore, the interstitial region comprises the volume
406
that has no solid surfaces. The intraparticle fluid velocity distribution is mostly slow flow and has
407
little overlap with the interstitial velocity distribution.
408
The interstitial flow region of the porous particle bed is geometrically similar to the solid-sphere
409
bed (of equivalent radius rh ). For this reason it was thought that the flow distributions might also be
410
similar. This is incorrect because the interstitial region in a bed of solid spheres is bounded by the
411
sphere surfaces, but for a bed of SPPs or FPPs the interstitial region is bounded by the (imaginary)
412
hull. The fluid gap between the hull and the sol surfaces is on the order of the sol radius, and this
413
gap allows flow at the shell surface to approach the average pore velocity.
414
3.4. Mass flow and flux through the hull
415
Flux through the porous particle hull is the component of fluid momentum density normal to the
416
surface, f = ρv · n, with units of kg/m2 s, where v is the velocity at a point on the surface and ρ is the
417
fluid density. Conservation of mass implies that the integral of flux over the hull surface S is zero Z
Z f dA /
S
dA = 0 S
17
(6)
418
The average hull flux magnitude, f¯h , is therefore defined as f¯h =
Z
Z |f |dA /
S
dA
(7)
S
419
where S is the hull surface of all particles in the packed bed. The average total mass flow into a single
420
particle hull is then defined as Z f¯h Fh = · dA 2N S
421
(8)
with units of kg/s and with N equal to the number of particles.
422
Hull flux magnitudes are, in general, less than the superficial flux magnitude given by ρqz . This is
423
because fluid velocities near the hull surface are less than the average pore velocity, and their vectors
424
are nearly tangent to the surface. For the random SPP packed bed, the hull flux is ≈ 9% of the
425
interstitial flux. For the random FPP packed bed, the hull flux is ≈ 10% of the interstitial flux and
426
both are in the direction of the pressure gradient.
427
Flow into the SPP is significant despite the relatively small flux magnitude. Approximately one
428
third of the bed volume flows into the particles in the time required to travel one SPP diameter
429
at the superficial velocity, which is approximately one millisecond for Resol = 1.00 × 10−3 . The
430
corresponding nondimensional expression is Fh 1 2rh Fbh = · · ρ L3 q
(9)
431
where the first term on the right is the volume flow rate into the hull, the second term is the reciprocal
432
bed volume and the third term is the characteristic time to travel one particle diameter. Fbh is therefore
433
the number of bed volumes passing into the SPP hulls during the time required to travel one SPP
434
diameter at the superficial velocity. These values are 0.34 and 0.37 for the random and BCC SPP
435
packings, respectively. Fh , Fbh and f¯h are tabulated in Table 5.
436
Although higher pressure is needed to drive flow to the standard Resol through the FPP bed and
437
the resulting mass flow is greater (Tables 3 and 5), the difference in mass flow is proportional to the
438
difference in pressure. For the same pressure gradient, mass flow rates into the SPP and FPP will be
439
the same and the superficial velocity will be greater in the SPP bed.
440
Higher pressure is also needed to drive flow through the BCC bed compared to the random bed.
441
The BCC packing is denser, and higher pressure is required to achieve the standard Resol . However,
18
442
the mass flow rate into the particles is actually slightly lower in the BCC bed than the random
443
packing, despite the elevated pressure. The explanation is that the mass flow rate into the particle is
444
proportional to the interstitial flux, which is the product of the interstitial porosity and pore velocity,
445
and this product is lower in the BCC packing. Packing
ch F
f¯h (kg/m2 s)
3.29 × 10−1 2.88 × 10−1 3.74 × 10−1 3.13 × 10−1
4.27 × 10−1 6.07 × 10−1 4.02 × 10−1 5.72 × 10−1
Fh (kg/s)
random SPP 1.26 × 10−11 random FPP 1.78 × 10−11 BCC SPP 1.17 × 10−11 BCC FPP 1.69 × 10−11
ch , and average flux magnitude, f¯h , through the particle hull area. Table 5: Mass flow, Fh , nondimensional mass flow, F −3 All cases are simulated at Resol = 1 × 10 .
446
3.5. Flux distribution on the particle hull surface
447
The flux distribution over the hull surface is a density function describing the probability of a
448
given flux value being observed on the surface. Flux distributions were calculated for each porous
449
particle and averaged into the overall flux distribution.
450
The overall flux distribution has a Laplacian shape as shown in Figure 6. It has a sharp mode
451
at fh = 0 and decreases exponentially with distance from the origin. Figure 6 also shows the flux
452
distributions for the two individual particles with the least and greatest flux magnitude. These differ
453
in amplitude near the origin but are otherwise similar. The flux magnitude of these two particles is
454
23% less and 22% greater than the average flux magnitude for all particles, respectively.
455
3.6. Volumetric flux in the direction of the pressure gradient
456
Although a significant volume of fluid crosses the porous particle hull, the interstitial region
457
carries the vast majority of fluid through the bed. Considering flow only in the direction of the
458
pressure gradient, the interstitial volumetric flux is ≈ 97% of the total. The volume flowing into the
459
particles and the volume flowing through the interstitial region are both large. This apparent paradox
460
emphasizes the fact that the two are fundamentally different measurements.
461
Total hull area scales with the bed volume, L3 , while the bed cross section scales with L2 . Fluid
462
mass flow is proportional to surface area in both cases, so particle hull flow scales with L3 and bed
463
flow scales with L2 . The ratio of hull flow to bed flow will therefore change with the bed dimensions,
464
even if the Reynolds number remains constant. 19
465
Intraparticle flux is defined as the volumetric flux through the particles in the direction of the
466
pressure gradient. It is the difference between flux through the entire bed and flux through the
467
interstitial region. Table 6 gives the volumetric flux in the direction of the pressure gradient for the
468
intraparticle region, the interstitial region and the entire bed. Each of the flux values is defined as
469
the volumetric flow through the cross-sectional area of the bed. Packing
Volumetric flux (m/s) intraparticle interstitial bed (total)
random SPP random FPP random sphere BCC SPP BCC FPP BCC sphere
1.21 × 10−4 1.82 × 10−4 0 1.34 × 10−4 1.96 × 10−4 0
4.66 × 10−3 5.87 × 10−3 3.13 × 10−3 4.36 × 10−3 5.67 × 10−3 2.78 × 10−3
4.70 × 10−3 6.06 × 10−3 3.13 × 10−3 4.49 × 10−3 5.86 × 10−3 2.78 × 10−3
Area (m2 ) 3.96 × 10−10 3.96 × 10−10 4.03 × 10−10 2.40 × 10−11 2.40 × 10−11 2.46 × 10−11
Table 6: Volumetric flux in the direction of the pressure gradient for the intraparticle region, interstitial region and bed. Flow in each bed has the same mean pore velocity, corresponding to Resol = 1 × 10−3 . Area is the cross-sectional area of the bed.
470
3.7. Internal fluid velocity
471
The pore velocity magnitude is the length of the velocity vector, ||v||, at a given point or averaged
472
over a given volume, and it is used to characterize flow within the porous particle as a function of
473
radial distance from its center. Figure 7 shows the average velocity magnitude as a function of radial
474
position inside the particle hull. The average magnitude is normalized by the overall mean pore
475
velocity, v z .
476
The average velocity magnitude at the SPP hull surfaces is ≈ 30% of the average velocity mag-
477
nitude in the entire bed volume. Moving from the hull toward the center, the average magnitude
478
decreases exponentially over a distance of one sol diameter and remains approximately constant over
479
the remaining radial distance to the core. The magnitude of this relatively constant velocity value is
480
approximately two orders of magnitude smaller than the average surface magnitude. Individual SPP
481
particles deviate from this profile, but the difference is small. The particle with the greatest internal
482
velocity magnitude has a profile that varies from 20% to 30% higher than the average profile. The
483
particle with least internal velocity magnitude varies from 20% to 35% below the average.
484
The FPP and SPP results are similar, observing that the region of diminished pore velocity in
485
the FPP spans nearly the entire particle radius. The mean pore velocity is slightly lower in the FPP, 20
486
as shown in Table 3, and the velocity range is slightly wider, as illustrated by the minimum and
487
maximum curves in Figures 7A and 7B. The extent to which a wider velocity range within the shells
488
of particles affects zone broadening in a chromatography column is not clear.
489
3.7.1. Reverse flow
490
In a pressure-driven flow through a packed bed of spheres, ring vortices may form at the down-
491
stream side of a sphere at very small Reynolds numbers. This phenomena is well known experimentally
492
in creeping flow past tangent spheres (Van Dyke and Van Dyke, 1982) and from a flow simulation in
493
a crystalline array of spheres (Maier et al., 2000). These vortices contribute a small amount of reverse
494
flow, which is perhaps best described as a local recirculation caused by contracting-then-expanding
495
flow. The phenomena is essentially absent in a random array of spheres, at least on a length scale
496
comparable to the pore size, because the random arrangement destroys the symmetry necessary for
497
such flow structures. Nevertheless, random sphere packs do have very small amounts of reverse flow.
498
However, these are to be found in pores oriented transversely to the pressure gradient, where the local
499
flow direction forms a very slight angle against the general flow.
500
In a BCC sphere packing, if solid spheres are replaced by porous particles, the sol structure
501
inhibits the development of pore-scale vortices. However, the interstitial region still has symmetrical
502
contracting and then expanding flow, and the resulting back pressure induces reverse flow in the
503
downstream side of the SPP sol, shown in Figure 8. This is in roughly the same location where the
504
ring vortex would appear between solid spheres. However, in the case of a random pack of SPPs, flow
505
against the gradient is resisted in the sol layers.
506
Also shown in Figure 8 are the FPP velocity field and probability density. While the FPP and
507
SPP interstitial flows are similar, the FPP velocity field is more collimated in the vortex region; the
508
reverse flow is more tightly focused, and its magnitude is less than in the vortex region of the SPP.
509
3.8. Perfusion in the model particle
510
In early work describing perfusion chromatography (Afeyan et al., 1991), an equation was devel-
511
oped (van Kreveld and van den Hoed, 1978) that equates the pore velocity to the increase in effective
512
pore diffusion coefficient Deff :
Deff = D + v · dp /2
(10)
513
where D is the solute pore diffusion coefficient in the absence of flow, dp is the particle diameter and
514
v is the velocity of fluid in the pore. To estimate Deff in a typical chromatography column, we assume 21
515
a mean pore velocity of v = 1.5 × 10−3 m s−1 . From Table 3 we note the ratio of mean perfusion
516
velocity (intraparticle) to overall mean pore velocity is 0.067 in a random SPP packing. We therefore
517
extrapolate the mean perfusion velocity in a typical chromatography column as v = 1.0 × 10−4 =
518
0.067 × 1.5 × 10−3 m s−1 . Inserting this into Equation 10, along with dp = 4.30 × 10−6 m, gives
519
Deff = D + 4.3 × 10−10 m2 s−1 or Deff = D + 4.3 × 10−6 cm2 s−1 .
520
This suggests that for small proteins where D ≈ 1 × 10−6 cm2 s−1 , Deff is enhanced by more than
521
a factor of four under these conditions. Because the solid phase mass transport efficiency is inversely
522
proportional to the effective pore diffusion coefficient (Giddings, 1965; Afeyan et al., 1991; Frey et al.,
523
1993; Carta and Jungbauer, 2010), the efficiency term increases thus lowering the solid phase plate
524
height by a factor of four. This is significant for biomolecule separations where diffusion coefficients
525
are small. Larger flow rates and interstitial velocities will increase v and cause an increase in Deff .
526
In the previous analysis, the value of D will typically be less than 1 × 10−6 cm2 s−1 , typical
527
of biomolecules. For example, the free solution diffusion coefficient of lysozyme with a molecular
528
weight of ≈ 14 kDa is ≈ 1 × 10−6 cm2 s−1 (Tanford, 1961; Mattisson et al., 2000). This suggests
529
that the pore diffusion coefficient will be smaller than this value (the pore diffusion coefficient is
530
always less than the free solution diffusion coefficient), and the effective pore diffusion coefficient will
531
be increased significantly by convection. The increase in convection-aided effective pore diffusion
532
becomes significant for molecules of higher molecular weight (and lower diffusion coefficients) than
533
lysozyme when the molecular size and weight is larger and all other conditions are held constant. When
534
the molecular size is a significant fraction of the pore size, restricted diffusion becomes important and
535
lowers chromatographic efficiency (Carta and Jungbauer, 2010; Wagner et al., 2017). Under the
536
conditions discussed here, the pore velocity may be large enough to help reduce this effect.
537
4. Discussion and Conclusions
538
Porous particles have much larger specific surface area than solid particles and higher pressures
539
are required to drive the flow to a specified pore Reynolds number. However, the pressure required to
540
achieve a specified volumetric flux is comparable for solid, SPP and FPP particles of equivalent hull
541
radius because the flux also depends on the bed volume. In fact, the SPP actually requires a slightly
542
lower pressure than solid spheres to achieve a specified superficial velocity. This result is reflected
543
in the fact that bed permeabilities hardly differ among the solid, SPP and FPP packed beds. FPP
544
permeability predicted by Brinkman’s equation is accurate, though the predicted SPP permeability
22
545
is less accurate due to the heterogeneous structure.
546
The SPP and FPP particles contain a significant fraction of the bed void volume, but the interstitial
547
void space carries the majority of fluid through the bed. The average fluid speed at the particle surfaces
548
is significant, but only a fraction of the flow is directed into the particle. The velocity magnitude
549
drops exponentially with internal distance from the surface. The average intraparticle pore velocity
550
within the shell is approximately two orders of magnitude less than the interstitial pore velocity.
551
Although the average intraparticle velocity is small in the direction of the pressure gradient,
552
volumetric flow into the particles is significant. Approximately one-third of the bed void volume flows
553
into the particle surfaces in the time required for fluid to travel one particle diameter at the superficial
554
velocity. Thus, the rate of fluid exchange is substantial near the particle surfaces.
555
One of the least understood aspects of LC is the convection and diffusion of large solute molecules
556
passing near, into and out of pores. For example, the rate of diffusion of proteins into and out of
557
pores limits the resolution of chromatographic separation due to the small diffusion coefficient which
558
is characteristic of larger molecules (Carta and Jungbauer, 2010). When the molecular size of the
559
solute is of the order of the pore diameter or characteristic length, even slower diffusion occurs into
560
and out of the pores. This effect is hydrodynamic in origin and is referred to as restricted diffusion
561
(Brenner and Gaydos, 1977; Dechadilok and Deen, 2006; Carta and Jungbauer, 2010).
562
Although the Brinkman level of approximation can be extended to the prediction of average
563
intraparticle velocities, direct simulation methods, such as the LB method used here, allow inspection
564
of the flow field at the pore level and, ultimately, the determination of effective solute diffusion as a
565
function of position in the porous particle. The authors are currently using this approach to model
566
the motion of finite-size particles driven by convection. This will hopefully facilitate the design of
567
new particles specifically for analytical bioseparations and may also explain the penetration depth
568
effect that appears in some adsorbents where the porous shell thickness seems to have a minor effect
569
on retention of medium to large biomolecules. A subsequent article will investigate aspects of mass
570
transport of finite-size solutes along with residence time and depth penetration in SPPs and FPPs.
571
Pore-level fluid simulation also can assist in the understanding of shear forces in polymer flows.
572
In size-exclusion chromatography (SEC) (Striegel et al., 2009b), degradation of polymers has long
573
been known to occur (Striegel, 2008; Striegel et al., 2009a,b), but the mechanisms of degradation are
574
inferred from related phenomena (Harrington and Zimm, 1965) and not directly observed. Under a
575
zero-concentration model of polymer flow, such that the presence of the polymer is assumed not to 23
576
affect the shear field, polymer extensional forces in the pore vicinity can be estimated using the present
577
methods. This would facilitate comparison with known physical data on bond scission (Caruso et al.,
578
2009).
579
In recent years, confocal laser scanning microscopy (CLSM) has provided particle-level details in
580
packed capillaries for particles with diameters greater than 1 µm (Reisinga et al., 2016). For sol
581
particles ≤ 100 nm in diameter, CLSM may not have the necessary resolution, so FIB and other
582
techniques for particle-level detail are required. Short of having a sol particle-resolved model for the
583
porous particle, packing methods, as used here, can be effective model builders for porous particles.
584
When model building software controls the shell thickness and particle defect generation, it is possible
585
to examine particle morphologies that not only model new experimental particles, but can be used as
586
particle models to explore new pore geometry schemes.
587
5. Symbol Glossary D
Solute pore diffusion coefficient
deff
Effective particle diameter
Deff
Solute effective pore diffusion coefficient
dp
Particle diameter
dsol
Sol particle diameter
f
Flux momentum density normal to particle surface
f¯h
Average hull flux magnitude
Fh
Mass flow into the hull
Fbh
Number of bed volumes passing into SPP hulls for finite time
k
Darcy (bed) permeability
kp
Intraparticle permeability
L
Length of cubic box used in velocity calculations
Ls
Porous shell thickness of SPP
n
Number of spheres per particle
N
Number of particles per bed
nsol
Number of sol particles per spherical particle
n
Surface normal
q, qz
Superficial velocity (Darcy flux)
24
r
radial distance
rc
Core radius
reff
Effective particle radius
rh
Hull radius
ri
Radius of the ith particle
rsol
Sol particle radius
Resol
Particle Reynolds number
Si
Surface area per particle
Sm
Mass specific surface area
SV
Surface area to volume ratio
v
Fluid velocity at a point
||v||
Euclidean norm of the vector v
vx , vy , vz
Velocity components in x, y and z
v
Mean pore velocity
vz
Mean velocity in the flow (z) direction
Vc
Core volume
Vh
Hull volume
Vp,m
Pore volume per mass of particles
Vs
Shell volume
Vt,s
Particle solid volume
∆x
Grid spacing
β
Nondimensional particle permeability
Interstitial porosity
p
Intraparticle porosity using the particle hull as the total volume
sp
Intraparticle porosity using the shell as the total volume
0p
The ratio of total pore volume to the total box volume
T
Total porosity
µ
Dynamic viscosity
ν
Kinematic viscosity
ρc
Density of the core 25
ρcs
Density of the core and shell
ρs
Density of the shell material
ρ
Density of fluid
ξi
Fraction of points used for surface area calculation
∇p, dp/dz Pressure gradient ∇v
588
Velocity spatial derivatives
Acknowledgments
589
The support of the National Institutes of Health under grant R44-GM108122-02 is gratefully
590
acknowledged. The authors acknowledge the Minnesota Supercomputing Institute (MSI) at the Uni-
591
versity of Minnesota for providing resources that contributed to the research results reported within
592
this paper. Additional thanks go to the W. M. Keck Center for Advanced Microscopy at the University
593
of Delaware for access to their electron microscope facility.
594
Supplementary material
595
Two files are given which have the coordinates to the synthetic sphere packs. These are 1030-27-
596
D-50-10-SPP and 1030-27-D-50-10-FPP shown in Table 1. The header information is contained in
597
the file header-info.
598
References
599
Afeyan, N., Fulton, S., Regnier, F., 1991. Perfusion chromatography packing materials for proteins
600
and peptides. J. Chromatogr. A 544, 267–279.
601
Allen, K. G., von Backstr¨om, T. W., Kr¨oger, D. G., 2013. Packed bed pressure drop dependence on
602
particle shape, size distribution, packing arrangement and roughness. Powder Techn. 246, 590–600.
603
Barrett, E. P., Joyner, L. G., Halenda, P. P., 1951. The determination of pore volume and area
604
distributions in porous substances. i. computations from nitrogen isotherms. J. Am. Chem. Soc. 73,
605
373–380.
606
Bear, J., 2014. Dynamics of fluids in porous media. Courier Dover Publications, New York.
26
607
608
Bijeljic, B., Muggeridge, A. H., Blunt, M. J., 2004. Pore-scale modeling of longitudinal dispersion. Water Resources Research 40 (11).
609
Brenner, H., Gaydos, L., 1977. The constrained Brownian movement of spherical particles in cylindri-
610
cal pores of comparable radius: models of the diffusive and convective transport of solute molecules
611
in membranes and porous media. J. Coll. Interf. Sci 58, 312–356.
612
613
614
615
616
617
Brinkman, H., 1949. On the permeability of media consisting of closely packed porous particles. Appl. Sci. Res. A 1 (1), 81–86. Brinkman, H. C., 1947. A calculation of the viscous force exerted by a flowing fluid on a dense swarm of particles. Appl. Sci. Res. A 1, 27–34. Brunauer, S., Emmett, P. H., Teller, E., 1938. Adsorption of gases in multimolecular layers. J. Am. Chem. Soc. 60, 309–319.
618
Carman, P. C., 1937. Fluid flow through granular beds. Trans. Inst. Chem. Eng. 15, 155–166.
619
Carta, G., Jungbauer, A., 2010. Protein Chromatography. Wiley-VCH, Weinheim, Germany.
620
Caruso, M. M., Davis, D. A., Shen, Q., Odom, S. A., Sottos, N. R., White, S. R., Moore, J. S., 2009.
621
622
623
Mechanically-induced chemical changes in polymeric materials. Chem. Rev. 109, 5755–5798. Crawford, C. W., Plumb, O. A., 1986. The influence of surface roughness on resistance to flow through packed beds. J. Fluids Eng. 108, 343–347.
624
Daneyko, A., Hlushkou, D., Baranau, V., Khirevich, S., SeidelMorgenstern, A., Tallarek, U., 2015.
625
Computational investigation of longitudinal diffusion, eddy dispersion, and transparticle mass trans-
626
fer in bulk, random packings of coreshell particles with varied shell thickness and shell diffusion
627
coefficient. J. Chromatogr. A 1407, 139–156.
628
629
630
631
de Neuville, B., Lamprou, A., Morbidelli, M., Soos, M., 2014. Perfusive ion-exchange chromatographic materials with high capacity. J. Chromatogr. A 1374, 180–188. Dechadilok, P., Deen, W., 2006. Hindrance factors for diffusion and convection in pores. Ind. Eng. Chem. Res. 45, 6953–6959.
27
632
DeStefano, J. J., Langlois, T. J., Kirkland, J. J., 2008. Characteristics of superficially-porous silica
633
particles for fast HPLC: Some performance comparisons with sub-2µm particles. J. Chrom. Sci. 46,
634
254–260.
635
636
637
638
639
640
Durlofsky, L., Brady, J. F., 1987. Analysis of the Brinkman equation as a model for flow in porous media. Phys. Fluids 30, 3329–3341. Eisfeld, B., Schnitzlein, K., 2001. The influence of confining walls on the pressure drop in packed beds. Chem. Eng. Sci. 56, 4321–4329. Frey, D. D., Schweinheim, E., Horv´ath, C., 1993. Effect of intraparticle convection on the chromatography of biomolecules. Biotechn. Progr. 9, 273–284.
641
Giddings, J. C., 1965. Dynamics of Chromatography. Marcel Dekker, New York.
642
Gritti, F., Cavazzini, A., Marchetti, N., Guiochon, G., 2007. Comparison between the efficiencies
643
of columns packed with fully and partially porous C18 -bonded silica materials. J. Chromatogr. A
644
1157 (1), 289–303.
645
646
647
648
649
650
651
652
653
654
Gropp, W., Lusk, E., Skjellum, A., 1999. Using MPI: Portable parallel programming with the messagepassing interface, 2nd Edition. The MIT Press, Cambridge. Harrington, R. E., Zimm, B. H., 1965. Degradation of polymers by controlled hydrodynamic shear. J. Phys. Chem. 69, 161–175. Hayes, R., Ahmed, A., Edge, T., Zhang, H., 2014. Core-shell particles: preparation, fundamentals and applications in high performance liquid chromatography. J. Chromatogr. A 1357, 36–52. Hlushkou, D., Khirevich, S., Apanasovich, V., Seidel-Morgenstern, A., Tallarek, U., 2007. Porescale dispersion in electrokinetic flow through a random sphere packing. Anal. Chem. 79, 113–121. Horv´ath, C., Lipsky, S. R., 1969. Column design in high pressure liquid chromatography. J. Chrom. Sci. 7, 109–116.
655
Iler, R. K., 1979. The chemistry of silica. John Wiley, Inc., New York.
656
Ismail, O. H., Pasti, L., Ciogli, A., Villani, C., Kocergin, J., Anderson, S., Gasparrini, F., Cavazzini,
657
A., Catani, M., 2016. Pirkle-type chiral stationary phase on core–shell and fully porous particles: 28
658
Are superficially porous particles always the better choice toward ultrafast high-performance enan-
659
tioseparations? J. Chromatogr. A 1466, 96–104.
660
Kandhai, D., Hlushkou, D., Hoekstra, A. G., Sloot, P. M., Van As, H., Tallarek, U., 2002. Influence
661
of stagnant zones on transient and asymptotic dispersion in macroscopically homogeneous porous
662
media. Physical review letters 88 (23), 234501.
663
664
665
666
667
668
669
670
671
672
673
674
675
676
677
678
Khirevich, S., 2011. High-performance computing of flow, diffusion, and hydrodynamic dispersion in random sphere packings. Ph.D. thesis, Philipps-Universit¨at Marburg Fachbereich Chemie. Kr¨ uger, T., Kusumaatmaja, H., Kuzmin, A., Shardt, O., Silva, G., Viggen, E. M., 2016. The Lattice Boltzmann Method: Principles and Practice. Springer, New York. Lloyd, L., Warner, F., 1990. Preparative high-performance liquid chromatography on a unique highspeed macroporous resin. J. Chromatogr. A 512, 365–376. Maier, R., Bernard, R., 2010. Lattice-Boltzmann accuracy in pore-scale flow simulation. J. Comp. Phys. 229 (2), 233–255. Maier, R. S., Kroll, D., Kutsovsky, Y., Davis, H., Bernard, R., 1998. Simulation of flow through bead packs using the lattice Boltzmann method. Phys. Fl. 10 (1), 60–74. Maier, R. S., Kroll, D. M., Bernard, R. S., Howington, S. E., Peters, J. F., Davis, H. T., 2000. Pore-scale simulation of dispersion. Phys. Fl. 12 (8), 2065–2079. Maier, R. S., Kroll, D. M., Bernard, R. S., Howington, S. E., Peters, J. F., Davis, H. T., 2003. Hydrodynamic dispersion in confined packed beds. Phys. Fl. 15 (12), 3795–3815. Masliyah, J. H., Neale, G., Malysa, K., van de Ven, T. G. M., 1987. Creeping flow over a composite sphere: solid core with porous shell. Chem. Eng. Sci. 42, 245–253.
679
Mattisson, C., Roger, P., Jonsson, B., Axelsson, A., Zacchi, G., 2000. Diffusion of lysozyme in gels
680
and liquids: A general approach for the determination of diffusion coefficients using holographic
681
laser interferometry. J. Chromatogr. B 743, 151–167.
682
683
MPI Forum, 2016. MPI: A Message-Passing Interface Standard, Version 3.1, June 4, 2015. http://mpi-forum.org/docs/mpi-3.1/mpi31-report.pdf (Nov. 2016). 29
684
685
686
687
688
689
690
691
692
693
694
695
Neale, G., Epstein, N., Nader, W., 1973. Creeping flow relative to permeable spheres. Chem. Eng. Sci. 28 (10), 1865–1874. Nemec, D., Levec, J., 2005. Flow through packed bed reactors: 1. single-phase flow. Chem. Eng. Sci. 60, 6947–6957. Neue, U. D., 1997. HPLC Columns: Theory, technology, and Practice. Wiley-VCH publishing, New York. Pfeiffer, J. F., Chen, J. C., Hsu, J. T., 1996. Permeability of gigaporous particles. AIChE J. 42, 932–939. Quinn, H. M., 2014. A reconciliation of packed bed permeability data: column permeability as a function of particle porosity. J. Mat. 2014, 1–22. Reimer, L., 2013. Scanning electron microscopy: Physics of image formation and microanalysis. Springer Publishing, New York.
696
Reisinga, A. E., Godinho, J. M., Hormanna, K., Jorgenson, J. W., Tallarek, U., 2016. Larger voids
697
in mechanically stable, loose packings of 1.3 m frictional, cohesive particles: Their reconstruction,
698
statistical analysis, and impact on separation efficiency. J. Chromatogr. A 1436, 118–132.
699
700
701
702
Rodrigues, A., Lu, Z., Loureiro, J., Carta, G., 1993. Peak resolution in linear chromatography: effects of intraparticle convection. J. Chromatogr. A 653, 189–198. Rodrigues, A. E., 1997. Permeable packings and perfusion chromatography in protein separation. J. Chromatogr. B 699, 47–61.
703
Schlueter, E., Witherspoon, P., 1994. Note on the validity of the Kozeny-Carman formulas for con-
704
solidated porous media. Tech. rep., Lawrence Berkeley Lab., CA (United States); California Univ.,
705
Berkeley, CA (United States).
706
707
Schure, M. R., Maier, R. S., 2006. How does column packing microstructure affect column efficiency in liquid chromatography? J. Chromatogr. A 1126 (1), 58–69.
708
Schure, M. R., Maier, R. S., Kroll, D. M., Davis, H. T., 2002. Simulation of packed-bed chromatogra-
709
phy utilizing high-resolution flow fields: comparison with models. Anal. Chem. 74 (23), 6006–6016. 30
710
711
712
713
Schure, M. R., Moran, R. E., 2017. Size exclusion chromatography with superficially porous particles. J. Chromatogr. A 1480, 11–19. Schuster, S., Wagner, B. M., Boyes, B. E., Kirkland, J. J., 2013. Optimized superficially porous particles for protein separations. J. Chromatogr. A 1315, 118–126.
714
Schuster, S. A., Boyes, B. E., Wagner, B. M., Kirkland, J. J., 2012. Fast high performance liquid
715
R chromatography separations for proteomic applications using fused-core silica particles. J. Chro-
716
matogr. A 1228, 232–241.
717
Song, C., Wang, P., Makse, H. A., 2008. A phase diagram for jammed matter. Nature 453, 629–632.
718
Spaid, M. A., Phelan Jr, F. R., 1997. Lattice Boltzmann methods for modeling microscale flow in
719
720
721
722
723
724
725
726
727
728
729
fibrous porous media. Phys. fluids 9 (9), 2468–2474. Sterling, J. D., Chen, S., 1996. Stability analysis of lattice Boltzmann methods. J. Comp. Phys. 123 (1), 196–206. Striegel, A. M., 2008. Observations regarding on-column, flow-induced degradation during SEC analysis. J. Liq. Chrom. Rel. Tech. 31, 3105–3114. Striegel, A. M., Isenberg, S. L., Cˆot`, G. L., 2009a. An SEC/MALS study of alternan degradation during size-exclusion chromatographic analysis. Anal. Bioanal. Chem. 394, 1887–1893. Striegel, A. M., Yau, W. W., Kirkland, J. J., Bly, D. D., 2009b. Modern Size-Exclsuion Liquid Chromatography, 2nd Edition. John Wiley, Inc., New York. Sullivan, S., Sani, F., Johns, M., Gladden, L., 2005. Simulation of packed bed reactors using lattice Boltzmann methods. Chemical Engineering Science 60 (12), 3405–3418.
730
Tanford, C., 1961. Physical chemistry of macromolecules. John Wiley, Inc., New York.
731
Tildesley, D., Allen, M., 1987. Computer simulation of liquids. Clarendon, Oxford.
732
Torquato, S., Truskett, T. M., Debenedetti, P. G., 2000. Is random close packing of spheres well
733
734
735
defined? Phys. Rev. Lett. 84, 2064–2067. Van Dyke, M., Van Dyke, M., 1982. An album of fluid motion. Vol. 176. Parabolic Press Stanford, p. 16-17. 31
736
737
van Kreveld, M. E., van den Hoed, N., 1978. Mass transfer phenomena in gel permeation chromatography. J. Chromatogr. 149, 71–91.
738
von Laven, K., 2015. Grid sphere. https://www.mathworks.com/matlabcentral/fileexchange/28842-grid-s
739
Wagner,
A.
J.,
2008.
A
practical
introduction
to
the
lattice
Boltzmann
method.
740
https://www.ndsu.edu/fileadmin/physics.ndsu.edu/Wagner/LBbook.pdf, Additional notes
741
for Statistical Mechanics 463/663 at NDSU.
742
743
Wagner, B. M., Schuster, S. A., Boyes, B. E., Kirkland, J. J., 2012. Superficially porous silica particles with wide pores for biomacromolecular separations. J. Chromatogr. A 1264, 22–30.
744
Wagner, B. M., Schuster, S. A., Shields, T. J., Boyes, B. E., Miles, W. L., Haynes, M. J., Moran,
745
R. E., Kirkland, J. J., Schure, M. R., 2017. Superficially porous particles with 1000˚ A pores for large
746
biomolecule HPLC and polymer SEC. J. Chromatogr. A 1489, 75–85.
747
White, F. M., Corfield, I., 2006. Viscous fluid flow. Vol. 3. McGraw-Hill New York.
748
Wu, Y., Abraham, D., Carta, G., 2015. Particle size effects on protein and virus-like particle adsorption
749
750
751
752
753
on perfusion chromatography media. J. Chromatogr. A 1375, 92–100. Yao, N. (Ed.), 2007. Focused ion beam systems: basics and applications. Cambridge U. Press, New York. Zick, A. A., Homsy, G. M., 1982. Stokes flow through periodic arrays of spheres. J. fluid mech. 115, 13–26.
32
Figure 1: Visualization of the experimental (A, B) and synthetic (C, D) SPP. A. SEM B. FIB cross-section C and D. Different external views showing the synthetic particle 1030-27-D-50-10-SPP. The blue color in D shows the second, deeper layer of the sol particles.
33
Figure 2: Local particle porosity vs. nondimensional radial distance from the particle center for SPP and FPP particles. The radial distance is made nondimensional by dividing r by the sol diameter dsol . The broken vertical line denotes the SPP core particle radius. The inset provides greater detail of the porosity in the outermost spherical shell.
34
Figure 3: A. Random packing of 125 spheres in a cubic bounding box with periodic boundaries and B. Random packing of 125 SPP in a cubic bounding box with periodic boundaries.
35
Figure 4: Main plot: Intraparticle and interstitial nearest-surface distributions for the random SPP packed bed. The probability density is plotted as a function of the nearest-surface distance normalized by the sol diameter. The sum of the two distributions is normalized to have unit area. Inset: FPP and SPP intraparticle distributions. The FPP distribution is normalized to unit area and the SPP distribution has area proportional to its pore volume relative to the FPP. The SPP intraparticle distributions (red curves) in the main plot and the inset are the same distributions with different normalizations.
36
Figure 5: Main plot: Velocity distributions in the FPP (solid line) and SPP (dashed line) random beds. P (v) denotes the relative volume of fluid with velocity v as a fraction of the bed pore volume. v/hvi denotes the z-component of fluid velocity normalized by the mean value, which is the same for FPP and SPP. Red, green and blue curves denote the intraparticle, interstitial and combined distributions. Inset plots show the full range of the interstitial and intraparticle distributions.
37
Figure 6: A. Hull flux distributions for random SPP packed bed. B. Distribution of surface flux momentum density 2 normal to the particle R surface (f = ρv · n) in the FPP and SPP beds (kg/m s). P r(f ) is the normalized distribution density, such that P r(f )df = 1.
38
Figure 7: Relative velocity as a function of depth in the particle for A. SPP and B. FPP. These results derive from a random packing of 125 spheres in a cubic bounding box with PBCs. In these figures, r is the radial distance from the center of the core, rh is the hull radius and dsol is the shell thickness.
39
Figure 8: Velocity in the direction of the pressure gradient (left to right) in BCC packed beds for A. SPP and B. FPP. The probability density is plotted against vz /v z . The inset shows contours of the velocity field and flow vectors in the x-z plane. Colors map to velocity as: red, vmin < v < −0.01 v; orange, −0.01 v < v < 0; black, 0; yellow, 0 < v < 0.01 v; blue, 0.01 v < v < 0.1 v; gray, 0.1 v < v < v; and green, v < v < vmax .
40
*Highlights
Highlights: Models of fully porous and superficially porous particles are constructed Extensive fluid flow is calculated within and around these particles in packed beds Velocities within these wide pore particles show significant internal flow Perfusion chromatography can be performed with these particles Course-grained theory agrees with the fluid velocities