You are on page 1of 37

Subscriber access provided by TULANE UNIVERSITY

Industrial & Engineering Chemistry Research is published by the American Chemical


Society. 1155 Sixteenth Street N.W., Washington, DC 20036
Published by American Chemical Society. Copyright American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the course
of their duties.
Article
EXPERIMENTALLY VALIDATED CFD SIMULATIONS OF MULTI-
COMPONENT HYDRODYNAMICS AND PHASE DISTRIBUTION
IN AGITATED HIGH SOLID FRACTION BINARY SUSPENSIONS
Li Liu, and Mostafa Barigou
Ind. Eng. Chem. Res., Just Accepted Manuscript DOI: 10.1021/ie3032586 Publication Date (Web): 26 Jul 2013
Downloaded from http://pubs.acs.org on September 1, 2013
Just Accepted
Just Accepted manuscripts have been peer-reviewed and accepted for publication. They are posted
online prior to technical editing, formatting for publication and author proofing. The American Chemical
Society provides Just Accepted as a free service to the research community to expedite the
dissemination of scientific material as soon as possible after acceptance. Just Accepted manuscripts
appear in full in PDF format accompanied by an HTML abstract. Just Accepted manuscripts have been
fully peer reviewed, but should not be considered the official version of record. They are accessible to all
readers and citable by the Digital Object Identifier (DOI). Just Accepted is an optional service offered
to authors. Therefore, the Just Accepted Web site may not include all articles that will be published
in the journal. After a manuscript is technically edited and formatted, it will be removed from the Just
Accepted Web site and published as an ASAP article. Note that technical editing may introduce minor
changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers
and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors
or consequences arising from the use of information contained in these Just Accepted manuscripts.

EXPERIMENTALLY VALIDATED CFD SIMULATIONS OF MULTI-
COMPONENT HYDRODYNAMICS AND PHASE DISTRIBUTION IN AGITATED
HIGH SOLID FRACTION BINARY SUSPENSIONS



L. Liu and M. Barigou*

School of Chemical Engineering, University of Birmingham, Edgbaston, Birmingham, B15 2TT, UK

Abstract
The mixing of dense (5-40 wt%) binary mixtures of glass particles in water has been studied in a stirred vessel
at the just-suspended speed and at speeds above it, using an Eulerian-Eulerian CFD model. For each phase
component, numerical predictions are compared to 3-D distributions of local velocity components and solid
concentration measured by an accurate technique of positron emission particle tracking. For the first time, it has
been possible to conduct such a detailed pointwise validation of a CFD model within opaque dense multi-
component slurries of this type. Predictions of flow number and mean velocity profiles of all phase components
are generally excellent. The spatial solids distribution is well predicted except near the base of the vessel and
underneath the agitator where it is largely overestimated; however, predictions improve significantly with
increasing solid concentration. Other phenomena and parameters such as particle slip velocities and
homogeneity of suspension are analysed.






Keywords: CFD; mixing; PEPT; stirred vessel; suspension








_______________________
*Corresponding author: Tel: +44 (0)121 414 5277 Fax: +44 (0)121 414 5324 Email: m.barigou@bham.ac.uk
Page 1 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
2

1. Introduction
The mixing of solids in liquids finds a wide range of industrial applications such as in the manufacture of fine
chemicals, pharmaceuticals, personal/home care products, paper and pulp, polymers, and food. These systems
give rise to immense difficulties and continue to pose a formidable challenge to industry as well as academic
research. There are huge practical difficulties in imaging solid-liquid flows and measuring local phase
velocities and phase concentrations, and problems are confounded by large particles with significant inertia,
wide particle size distributions and high solid fractions. Consequently, the methods generally adopted for
designing stirred vessels for solid-liquid mixing tend to be based on a global black-box approach, and a more
detailed description of the internal flow structure has long been missing to enable the development of more
rational rules for establishing process parameters and equipment design. The advent of powerful measurement
and modelling techniques in recent years has reenergised interest in this field and more effort is increasingly
being devoted to fully understand the mechanisms behind particle suspension and distribution
1-4
.

Among the various flow visualisation techniques, laser Doppler velocimetry (LDV) and particle image
velocimetry (PIV) have become the most reliable to examine the complex nature of the flow fields in optically
transparent single-phase systems
5-7
. These techniques, however, fail completely in dense systems which are
opaque and their application to multiphase flows has, therefore, been restricted to very dilute suspensions
1
. One
of the important aspects in the local description of multiphase mixtures is the measurement of phase distribution
which has remained a challenge. Attempts at local measurements in these systems have been mainly limited to
the measurement of mean axial solid-concentration profiles at relatively low concentrations, using a single
vertical conductivity or capacitance probe traverse or a withdrawal technique
8,9
. More recent experimental
studies using electrical resistance tomography (ERT) demonstrated, albeit mainly qualitatively, how
visualisation of gas, solid or liquid distribution can help improve understanding of mixing processes
10
.
However, none of these methods is suitable to probe concentrated suspensions in detail to acquire quantitative
information on the local flow dynamics of the different phase components or their individual 3D distribution.

Rammohan et al.
11
developed a more sophisticated computer automated radioactive particle tracking (CARPT)
technique based on gamma-ray emissions. It has been argued for some time that Lagrangian data obtained from
a single particle trajectory, where such data can be accurately determined, may unravel valuable mixing
information which is not provided by Eulerian observations
12-17
. We recently reported in a series of papers on
the successful use of the technique of positron emission particle tracking (PEPT) to measure the local
hydrodynamics as well as the spatial distribution of each phase component in mechanically agitated solid-liquid
suspensions
16,18-20
.

On the computational front, there have been a number of studies on solid-liquid suspensions ranging from dilute
to concentrated systems. Direct or Large Eddy Simulations are still far too expensive computationally to deal
with multiphase flows in stirred tanks and can only handle dilute systems (~ 5 vol%). Therefore, solution of the
Navier-Stokes equations by means of Computational Fluid Dynamics (CFD) codes with the appropriate two-
phase models incorporated seems to be the most advanced cost-effective method currently available to model
such systems.
Page 2 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
3


In most CFD studies of dilute solid-liquid suspensions ( ~ < 3 wt%), predictions of the phase velocities have
been compared to experimental data obtained from optical techniques such as LDV and DPIV published in the
work of Oshinowo and Bakker
21
and Fan et al.
22
. On the other hand, the CARPT technique
11
was used to obtain
the flow field of the solid phase for CFD validation at a solid concentration of only 2.1 wt%
23
. CFD predictions
of dilute solid distributions (up to ~ 1.33 wt%) have also been compared to experimental data obtained from an
optical attenuation technique, though measurements were limited to the axial solid distribution.
24-26


A number of numerical studies have also addressed suspensions with higher solid loadings up to ~ 34 wt%. The
validation work reported in the literature has been very limited, however, relying mostly on single-traverse
probe measurements of axial solid distribution
2,21,27
but no detailed hydrodynamic data such as the distribution
of phase velocities or slip velocities. It is only recently that the axial velocities of both the liquid and solid
phase measured by an ultrasound velocity profiler have been used to validate CFD predictions for medium
concentrated solid suspensions (13.6 wt%)
28
. The effects of different CFD approaches such as MFR (multi-
frame reference) and SG (Sliding Grid), as well as different drag force models on the prediction of the solid
distribution have also been studied.
26,27,29
However, the lack of detailed hydrodynamic data for validation has
limited the development of reliable approaches and models.

Overall, therefore, the accuracy of CFD in the mixing of dense suspensions has not been sufficiently studied
and, in particular, little information exists on flow hydrodynamics. Hence, a detailed validation using detailed
pointwise measurements of 3-D velocities as well as the distribution of the phases and phase components,
especially under conditions of moderate to high solid loadings, is needed in order to assess the capability of
CFD in these complex flows. The availability of the PEPT technique has provided an exceptional opportunity
for studying such systems and for using the unique measurements obtained to rigorously assess with
unprecedented detail and accuracy the capability of the numerical models and codes available.

In this paper, the flow of binary mixtures of coarse glass particles containing up to 40 wt% solids in water, is
studied numerically by means of a CFD model in a mechanically agitated vessel under different regimes of
particle suspension. Experimental results obtained using PEPT in our previous work
19
at N = N
js
together with
new measurements at N>N
js
are exploited for validation of a number of suspension characteristics, including
most importantly the detailed pointwise 3-D velocity field of each phase component and its spatial phase
distribution. The results are also analysed to infer for each phase component the local time-average particle-
fluid slip velocity, phase flow number and suspension uniformity. Finally, verification of the solids mass
balance and mass continuity in the vessel is presented.

2. Experimental
2.1 Positron emission particle tracking
The two-phase flow field and spatial phase distribution were determined by the technique of positron emission
particle tracking (PEPT). PEPT is a non-intrusive technique which can be applied to opaque systems. It uses a
single positron-emitting particle as a flow tracer which is then tracked in 3-D space and time to reveal its full
Page 3 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
4

Lagrangian trajectory. The measurements can be analysed in two different ways: a Lagrangian-statistical
analysis exploiting concepts such as residence time, circulation time and trajectory length distribution; or, a
Lagrangian-Eulerian analysis used to extract local Eulerian quantities from the purely Lagrangian information
contained in the tracer trajectory such as the three velocity components (
z
u ,
r
u ,

u ) of each phase component,


its local occupancy or time-average concentration
20
. PEPT has been validated in water by comparing its
measurements with PIV and has been shown to be an accurate and reliable technique
30
. More details about the
technique can be found in our earlier papers
14,16,18-20,31,32
.
2.2 Apparatus and procedures
The PEPT experiments were conducted in a fully-baffled flat-base Perspex vessel of diameter T = 288 mm,
agitated by a down-pumping 6-blade 45 pitched-turbine (PBT) of diameter D = 0.5T, as illustrated in Fig. 1.
The height of the suspension was set at H = T and the impeller off-bottom clearance was 0.25T. The suspending
liquid was an aqueous solution of NaCl of density 1150 kg

m
-3
NaCl was added to enable the water density to
be matched to that of the resin PEPT tracer used to track the liquid and the solid particles used were spherical
glass beads of density 2485 kg

m
-3
. Two nearly-monomodal particle size fractions, (d
p
~ 1 mm; ~ 3 mm) were
used to make binary solid-liquid suspensions. The two size fractions were mixed in equal proportions with a
total solid mass concentration, X, varying from 0 to 40 wt%, i.e., X
1
= X
2
= 0.5 X. Experiments were conducted
under different regimes of solid suspension corresponding to: (i) N = N
js
, the minimum speed for particle
suspension
19
; and (ii) speeds above N
js
up to 2N
js
. N
js
was visually determined in the transparent vessel
according to the well-known Zwietering criterion
33
, i.e., no particle remains stationary on the bottom of the tank
for longer than 1-2 s. The conditions of the PEPT experiments and CFD simulations conducted are summarized
in Table 1.

Each component of the suspension was individually tracked and its full trajectory separately determined in three
successive experiments. A neutrally-buoyant radioactive resin tracer of 600 m diameter was used to track the
liquid phase, and from each particle size fraction a representative glass particle was taken, was directly
irradiated via a cyclotron and its motion tracked within the agitated slurry. The PEPT tracking time was 30 min
in each case, long enough so that adequate data resolution was obtained in every region of the vessel and
ergodicity of flow, the theoretical condition which guarantees that a flow follower is representative of all the
solid or liquid phase it represents, could be safely assumed. Confirmation of the ergodicity assumption through
experimental verifications has been reported in previous papers
18,19,34


3. Numerical modelling
Numerical simulations were performed using the commercial CFD code ANSYS-CFX 12.0. In simulating
solid-liquid flows, the Eulerian-Lagrangian model would be a more realistic approach because it simulates the
solid phase as a discrete phase and so allows particle tracking. However, the number of particles that can be
tracked is currently very limited, thus restricting the applicability of the model to dilute mixtures. The multi-
fluid Eulerian-Eulerian model was therefore used instead, whereby the liquid and solid phases are both treated
as continua.

Page 4 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
5

The governing equations which form the basis of the model are the continuity equation:

( ) ( )
0 =

i
q iq q q q
x
u
t

(1)

and the momentum equation

( ) ( )
( )
s B q vm q lift g q q iq ij
k
kq
i
jq
j
iq
qeff q
i i
q
i
q jq iq q q iq q
P F F F F F
x
u
x
u
x
u
x x
P
x
u u
t
u
+ + + + +
(
(

|
|

\
|

, ,
3
2


(2)

where is density, is volumetric fraction, u is velocity vector, t is time, P is pressure,
eff
is effective
viscosity,
ij
is the Kronecker delta ( 1 =
ij
when i = j; 0 =
ij
when i j), and i, j, k are the three coordinate
directions; note that q d denotes the dispersed phase, and q c denotes the continuous phase
35
. The
gravitational force is denoted by F
g
and the inter-phase drag force is denoted by F
iq
. The lift force, F
lift,q
, and

the
virtual mass force, F
vm,q
, were initially included but then subsequently omitted as they did not affect the results,
a finding also previously reported by others
36, 37
. The term F
B
is the body force which includes the Coriolis and
centrifugal forces induced by using the Multiple Frames Reference (MRF) approach, as discussed further below.

Inclusion of the particle-particle interaction force for the solid phase in equation (2) was attempted through the
solid pressure term (
s
P ). As inter-particle interactions increase with solid concentration, the inclusion of this
term has been suggested particularly for highly concentrated suspensions (C

> 0.2). This solid pressure term is,
therefore, a function of the solid concentration
38
, thus

P
s
= P
s
(C
s
) (3)

and hence,

s s s
C C G P = ) ( (4)

The function G (C
s
) is referred to as the Elasticity Modulus, and is expressed as follows

) C C ( E
s
sm s
e G ) C ( G

=
0
(5)

where G
0
is the reference elasticity modulus, E is the compaction modulus, and C
sm
is the maximum packing
parameter (maximum solid loading). There are no generally accepted values for these parameters; however, the
values G
0
= 1 Pa, E = 20 600, have been suggested by Bouillard et al.
39
. The maximum packing parameter C
sm

was determined by Thomas
40
as 0.625 for spherical particles.

Page 5 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
6

The inclusion of particle-particle interactions through the solid pressure model which is available in ANSYS-
CFX, however, led to bad convergence due to linearisation problems in the software for all solid concentrations
investigated, and it was not possible to achieve convergence with residual levels below 10
-3
. In fact, better
predictions were obtained without inclusion of the solid pressure model which was consequently removed from
the simulations.

The code uses the finite-volume method to discretise the Navier Stokes equations. The so-called High
Resolution Advection Scheme was used to discretise the governing equations. The simulations were run in
steady-state, so the transient terms were also omitted. Note that for simplicity, the equations in this section are
written for one component only of the dispersed phase, but for a binary suspension they can be readily extended
to include a second component.

The inter-phase drag force, F
iq
, was modelled via the drag coefficient, C
D
, as:

( )
c d c d D
p
d d
iq
u u u u C
d
F =

4
3
(6)

C
D
was estimated using the Gidaspow drag model for densely distributed solid particles
38
:

for 2 . 0
d
( ) |

\
|
+ =

44 . 0 , Re' 15 . 0 1
Re'
24
max
687 . 0 65 . 1
c D
C (7)
and 2 . 0 >
d

( ) ( )
c d
p
c c
p c
c c
iq
u u
d d
F

1
4
7 1
150
2
2
(8)

where the particle Reynolds number Re is defined as
c
p c d c
c
d u u


= Re' , and
c
is the viscosity of the
continuous phase. The model has been found to be particularly useful for higher solid loadings in pipes
41, 42
and
in stirred vessels
26
.

The above system of differential equations (1) and (2) was closed by employing the well-known mixture k
turbulence model where the two phases are assumed to share the same k and
27
. The two-equation turbulence
model consists of the differential transport equation (9) for the turbulent kinetic energy, k, and equation (10) for
its dissipation rate, , as follows:

( ) ( )
( )



+
|
|

\
|

|
|

\
|
+

k q q
i k
qeff
q q
i i
iq q q q q
P
x
k
x x
k u
t
k
(9)

( ) ( )
( )

2 1
C P C
k x x x
u
t
k q q
i
qeff
q q
i i
iq q q q q
+
|
|

\
|

|
|

\
|
+

(10)
Page 6 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
7


where
k
P is a generation term of turbulence; 44 . 1
1
= C , 92 . 1
2
= C , 1 =
k
, and 314 . 1 =

.

The MFR method was used to take into account the interaction between the stationary baffles and the rotating
impeller blades. Consequently, the computational domain was divided into two regions, as shown in Fig. 1: one
domain containing all of the rotating elements (hub, blades and shaft), and the other containing the stationary
parts (baffles, tank wall and base) separated by three interfaces, a vertical cylindrical interface at r = 94 mm, and
two plane horizontal interfaces at z = 37 mm and 107 mm from the bottom of the tank. The interfaces were set
sufficiently remote from the impeller to minimize the effects on the results.

A mesh independence study was conducted. There were no significant improvements in either the velocity
profiles or the impeller torque beyond a mesh consisting of 1,065858 cells. However, as the computational cost
was not prohibitive, an even finer mesh was used to achieve a good prediction of turbulent quantities; this aspect
has been discussed by Deglon and Meyer
43
and Coroneo et al.
44
. Thus, the computational grid used consisted of
~ 1,504,928 non-uniformly distributed unstructured tetrahedral cells, as illustrated in Fig. 1. Inflated boundary
layers were used on the bottom of the tank, walls, baffles, blades and shaft to cover the high velocity gradients
in these regions of no slip. Numerical convergence was deemed to have occurred when the sum of all
normalized residuals fell below 10
-4
for all equations; however, most residuals were in fact usually below 10
-6
by
the end of the simulation.

4. Results and discussion
A detailed quantitative comparison between the CFD predictions and PEPT measurements was conducted at the
positions depicted in Fig. 1. For the liquid phase and both components of the solid phase (i.e., 1 mm and 3 mm
particles), the azimuthally-averaged distributions of the local velocity components (
z
u ,
r
u ,

u ) were
compared:

(i) axially along two radial positions: r = 0.53R, close to the tip of the impeller; and r = 0.90R, half a baffle away
from the tank wall; and

(ii) radially along five axial positions: z = 0.016H, close to the base of the vessel; z = 0.08H, approximately
halfway between the base and the lower edge of the impeller; z = 0.17H, just under the lower edge of the
impeller; z = 0.33H, just above the upper edge of the impeller; and z = 0.83H, in the upper part of the vessel.
For ease of comparison, all the velocity plots presented have been normalized by the impeller tip speed
( DN u
tip
= ).

The azimuthally-averaged spatial distributions of the local solid concentration were also resolved both by PEPT
and CFD, and compared at all these radial and axial locations. The solid concentration profiles presented have
been normalized by the overall mean volume solid concentration (C).

Page 7 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
8

4.1 Velocity distributions
4.1.1 Distribution of local velocity components
Azimuthally-averaged axial and radial distributions of the local velocity components (
z
u ,
r
u ,

u ) measured at
N
js
experimentally by PEPT, and the equivalent distributions predicted by the CFD model are shown in Fig. 2
and Fig. 3, respectively, for the liquid phase and the two solid components (d
p
= 1mm; 3 mm) at a solid
concentration of 5 wt%. The mixture components exhibit similar velocity distributions where
z
u is the
predominant velocity component, which is consistent with the performance of a down-pumping mixed flow
impeller. The radial distributions exhibit sharp variations in the impeller region (z = 0.17H, 0.33H), but are
fairly flat in the upper parts of the vessel (z = 0.83H). Near the tank base,
r
u is predominant as each component
of the mixture is forced by the pumping action of the impeller to move towards the wall and is entrained in the
single recirculation flow loop, thus, generated.

For the liquid phase and the 1 mm solids, all three velocity components are very well predicted by CFD
throughout the vessel. For the 3 mm particles, however, there are some significant discrepancies between CFD
and PEPT concerning the axial component
z
u . It is interesting to note that the number density of the 1 mm
particles in the suspension is 27 times that of the 3 mm particles. It is perhaps plausible that as the number
density of the solids increases, the assumption of a continuum inherent in the Eulerian-Eulerian model becomes
somewhat less unrealistic, which might explain why the accuracy of prediction is better for the smaller 1 mm
particles.

The numerically predicted axial velocity distributions at N
js
for solid concentrations of 10 wt%, 20 wt% and 40
wt%, are compared to the experimental distributions in Figures 4, 5 and 6, respectively. The distributions are
similar to those observed at 5 wt% solids, as discussed in the next section.

With increasing solid concentration, the accuracy of prediction for the liquid phase and the 1 mm solids remains
high. It is interesting, however, that for the 3 mm particles the numerical prediction improves significantly with
X, becoming very good at X = 20 wt% and X = 40 wt%. Again, as X increases, the solid phase tends to behave
increasingly more like a continuum, thus, possibly improving the suitability of the Eulerian-Eulerian model.
Furthermore, the Gidaspow drag model employed in this study was previously found to be particularly useful at
higher solid loadings, which may have perhaps contributed the most to the improved accuracy of prediction at
the higher solid concentrations
26, 41, 42
. On the other hand, however, as solid concentration increases, particle-
particle collisions are expected to increase which tends to complicate the physics of the flow to an unknown
extent and, hence, it is difficult at present to suggest a completely satisfactory explanation.

Numerical simulations were also conducted to assess the capability of CFD to predict the solid suspension at
N >> N
js
. Sample results are shown in Fig. 7 for 10 wt% solids at N = 1.8N
js
. Apart from some local
discrepancies affecting especially the particles around the impeller discharge stream, the flow field of each
mixture component is well predicted. At such a high rotational speed, the intensity of turbulence is extremely
high, and it is well known that the standard k- model used here tends to significantly underestimate the
Page 8 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
9

turbulence kinetic energy and dissipation rate, especially near the tip of the impeller
45, 46
; this may explain the
disparities observed near the impeller discharge. Similar results were obtained for the other solid
concentrations.

4.1.2 Reynolds number similarity
The flow structure at high Reynolds numbers presents similar characteristics, a phenomenon often referred to as
Reynolds number similarity. In fully turbulent flow, viscosity effects vanish and dimensionless mean values
such as velocities normalised by the impeller tip speed should be independent of Reynolds number. This
similarity has been reported in Newtonian and non-Newtonian liquids but does not seem to have been tested in
multiphase flows of the type studied here
47, 48
. Results plotted in Fig. 8 show the azimuthally-averaged radial
distributions of the normalised velocity components of the three mixture components at the four solid
concentrations investigated. The values of the impeller Reynolds number (Re
imp
= ND
2
/) are given in Table 1,
and cover the range 1.510
5
2.410
5
for N = N
js
. For the liquid phase, the Reynolds number similarity holds at
all solid concentrations. For the solid components, some minor disparities are observed but the principle of
similarity seems to hold too.

4.2 Spatial distribution of solid components
4.2.1 Method of evaluation of local phase concentration
In addition to location and velocity information, PEPT allows the particle occupancy distribution within the
vessel to be determined. By dividing the vessel volume into a grid of n
c
cells, occupancy can be obtained by
calculating the fraction of the total experimental time, t

, spent by the tracer in each cell during the experiment.


Such a definition establishes a mathematical identity between occupancy and probability of presence of the
particle tracer, but undesirably makes occupancy highly dependent on the density of the grid so that as the
number of cells increases occupancy tends to zero. If the cells are chosen to have equal volume, however, this
problem is circumvented by using the ergodic time defined as
c E
n t t /

= , instead of the total experimental


time t

.

The ergodic time represents the time that the tracer would spend in any cell if the flow were ergodic, an
asymptotic status in which the flow tracer has exactly an equal probability of presence everywhere within the
system. Thus, the local occupancy, O
j
, can be defined as:

E
j
j
t
t
O

= (11)
where t
j
is the time that the tracer of the j-th component (solid or liquid) of the mixture spends inside a given
cell. For the binary systems examined in this work, the index j assumes values 1, 2 or 3 where j = 1, 2 denotes
the solid components, and j = 3 denotes the liquid phase also designated by the symbol L.

On the basis of such a definition, it can be shown that, in a multiphase system, occupancy, O
j
, of component j is
directly related to its local phase volume concentration, c
j
, via the relationship
Page 9 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
10


j
j
j
C
c
O = (12)

where C
j
is the mean phase volume concentration of component j in the vessel. Full details of the mathematical
treatment can be found in our previous papers
18, 19


Thus, using equation (12) the spatial distribution of the local concentration of each flow component was fully
resolved in the binary suspensions investigated.

4.2.2 Comparison of experimental and CFD-predicted phase distributions
The ability to predict the spatial phase distribution is of the utmost importance in modelling suspensions in
stirred vessels because it impacts directly on the prediction of other important phenomena such as heat and mass
transfer. The azimuthally-averaged axial and radial distributions of the two solid components at the just-
suspended state corresponding to N
js
are displayed in Figs. 9 and 10, respectively. Overall, the agreement with
experiment is good, except near the base of the vessel and around the impeller. The solid distributions are
largely overestimated near the base (z = 0.016H), and likewise they are largely underestimated just beneath and
above the impeller (z = 0.17H, 0.33H); the errors are worse for the 3 mm particles. The worst CFD predictions
occur at the lowest solid concentration of 5 wt%, but they improve significantly both axially and radially with an
increase in solids loading, especially close to the tank base.

Particle lift-off from a bed of particles on the bottom of the vessel occurs as a result of the drag and lift forces
exerted by the moving fluid. The flow near the base has been described as boundary layer flow which causes
particles to be swept across the base of the vessel
49
. Once small fillets of particles have been formed, particle
lift-off is usually seen to be caused by sudden turbulent bursts originating in the turbulent bulk flow above.
Fluid-particle interactions are dependent on the relative size of eddies and the particles. Eddies relatively large
compared to the particles will tend to entrain them and are responsible for their suspension. Consequently,
different agitator designs and configurations generate different convective flows and, thus, achieve different
levels of solids suspension at the same power input. The complex nature of the flow field in a stirred vessel is
such that there is no fundamental theory to describe the process of particle suspension. The CFD model does
not incorporate all of the complex physics of particle sedimentation, particle lift-off, and particle-particle
interactions and, hence, the difficulty in achieving accurate prediction of the local solid concentration near the
base of the vessel.

Considering the radial solid distributions at N
js
shown in Fig. 10, the solids are nearly uniformly distributed
except in the central region underneath the impeller where there is a substantial mound of accumulated solids.
The axial solid distributions at N
js
depicted in Fig. 9, also indicate a fairly uniform suspension in the upper part
of the vessel, with the solid concentration reducing gradually upwards until it reaches zero close to the free
surface. However, large concentration gradients are found in the lower part of the vessel (z < 0.2H). As
Page 10 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
11

discussed in more detail later, the homogeneity of the solid distribution at N
js
improves when solids loading
increases, and this feature is captured by CFD.

At impeller speeds above N
js
, the uniformity of the suspension improves significantly, as shown in Fig. 11. The
agreement between CFD predictions and PEPT measurements improves somewhat, but predictions remain
rather poor near the base.

4.3 Uniformity of suspension
4.3.1 Binary suspensions
The uniformity of suspension can be estimated by the uniformity index, , introduced in our recent works
16,18
,
thus:

1
1
1
1
1
1
2 2
+ |

\
|
=
+
=

=
c
n
i
i
c
C
C c
n

(13)

where n
c
, as defined above, is the total number of cells in the PEPT measurement (or CFD computational) grid
and i is the cell number. The uniformity of the suspension is at its minimum when 0 (condition achieved
for a theoretically infinite variance,
2
, of the normalised local concentration); and the solids are uniformly
distributed within the vessel volume when = 1, i.e., the local solid concentration is equal to the average
concentration everywhere in the vessel (
2
= 0).

Results for the binary suspensions are plotted in Fig. 12. Generally, the 1 mm particles are better distributed
than the 3 mm particles at all speeds except when N >> N
js
. Initially rises sharply as N increases above N
js
, but
a plateau is reached at high speeds and it becomes difficult to exceed ~ 0.9, as it is hard to achieve perfect
homogeneity below the impeller and near the free surface. This state of dispersion is reached at speeds
significantly lower for the lighter 1 mm particles (~ 1.5N
js
) than for the 3 mm particles (~ 2N
js
).

It can also be inferred from the data in Fig. 12 that, for a given particle size, increases as a function of mean
solid concentration when the suspensions are considered at the same hydrodynamic mixing regime, i.e., at their
N
js
speed or the same multiple of their respective N
js
value. Note that the N
js
values for the different solid
concentrations investigated are displayed in Table 1. This seems to suggest, perhaps counter-intuitively, that for
a given state of suspension more concentrated suspensions will tend towards a homogeneous state simply by
virtue of their increased solids loading (C 1).

4.3.2 Comparison of binary and mono-disperse suspensions
We recently studied experimentally the suspension of the same particles considered here but in a mono-disperse
mode
18
. The spatial distributions of these particles obtained from PEPT measurements in the binary and mono-
disperse modes are compared in Fig. 13. The distributions of the 1 mm particles do not show any significant
difference at any solid concentration. The 3 mm particles, however, appear to be better distributed in the binary
Page 11 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
12

mixture than in the mono mixture at low to medium concentrations (5 wt% and 10 wt%), but this difference
disappears at higher concentrations. Such an observation is at present difficult to explain. Compared to the
binary suspension, however, significantly higher local solid concentrations are observed in the lower part of the
vessel when the 3 mm particles are in a mono-disperse mode, as shown in Fig. 13. This maldistribution is
confirmed by the uniformity index plots in Fig. 12 where the values of for the binary suspension are
significantly greater than for the mono suspension at the lower solid concentrations investigated.

4.4 Impeller pumping capacity
Recently, we showed how the usual definition of the flow number used to estimate the pumping effectiveness of
impellers in single-phase systems can be extended to a two-phase problem
18
. For each component j of the
mixture, Q
(j)
is calculated along the lower horizontal edge of the impeller blade corresponding to the impeller
discharge plane, by integrating the axial velocity profile of the component weighted by its local volume
concentration, i.e. by c
1
and c
2
for the solid components and by (1-c
1
-c
2
) for the liquid phase. Thus, the flow
number is computed using:

for component 1 of the solid phase (d
p
= 1 mm):

( ) ( )

= =
PBT
S
z
S
S
dS u c
ND ND
Q
Fl
1
1
3 3
) 1 (
1
1
(14)

for component 2 of the solid phase (d
p
=3 mm):

( ) ( )

= =
PBT
S
z
S
S
dS u c
ND ND
Q
Fl
2
2
3 3
) 2 (
2
1
(15)

and for the liquid phase:

( )
( )
( )
( )

= =
PBT
L
z
L
L
dS u c c
ND ND
Q
Fl
2 1
3 3
1
1
(16)

The sum of Q
(L)
, Q
(S1)
and Q
(S2)
represents the total volumetric discharge, Q, and introducing the total two-phase
flow number, Fl, it follows that:

( ) ( ) ) 2 ( 1
3
S S L
Fl Fl Fl
ND
Q
Fl + + = = (17)

The computational CFD results are presented in Table 2 alongside the experimental PEPT values for ease of
comparison.

Page 12 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
13

The solid flow numbers
) 1 (S
Fl

and
) 2 (S
Fl are only a small fraction of the liquid flow number Fl
(L)
, but they
increase proportionally to C. This is of course expected because in the calculation of the flow number, the
flowrate of each phase (component) is weighted by its local volume concentration (see Eqs. (14), (15) and (16)).
The experimentally measured values of
) 1 (S
Fl and
) 2 (S
Fl are approximately equal under all conditions
investigated, meaning that both solid components (1 mm and 3 mm) are non-preferentially pumped by the PBT.
The magnitude of
) (L
Fl reduces with increasing solid concentration, and so does the overall flow number Fl
because of the higher proportion of heavy solids which are being pumped.

At low solid concentrations and low agitation speeds,
) 1 (S
Fl and
) 2 (S
Fl are largely underpredicted by CFD
probably because of their small values which are more prone to numerical error. However, the predictions
improve greatly at higher mean solid concentrations or higher impeller speeds. The predictions of
) (L
Fl are
very good throughout, and because of its dominance, the overall flow number Fl is also well predicted with an
average error on the order of ~ 10%, as shown in Table 2 in Supporting Information.

There are no published flow number data in solid-liquid suspensions, but in the absence of particles (X = 0) the
single-phase flow number obtained from PEPT ( 87 . 0
) (
=
L
Fl ) agrees very well with published data for a 6-
blade pitched turbine (e.g., 86 . 0
) (
=
L
Fl in Gabriele et al.
50
; and 88 . 0
) (
=
L
Fl

in Hockey et al.
51
.

4.5 Distributions of particle-fluid slip velocities
Information on particle-fluid slip velocity is of real value to processing applications involving the transfer of
heat or mass, such as in chemical reactions or the sterilisation of particulate food mixtures. A crude assumption
often used in practice takes the free terminal settling velocity of the particle as a representative measure of its
mean slip velocity
52
. In 3-D the magnitude of the local time-average slip velocity vector of a solid component
can be estimated using PEPT (or CFD) data, thus:

( ) ( ) ( )
2
) ( ) (
2
) ( ) (
2
) ( ) ( S
z
L
z
S
r
L
r
S L
s
u u u u u u u + + =

(18)

where, as in the above, the superscripts (L) and (S) refer to the liquid and solid phase, respectively.

Azimuthally-averaged distributions of the local time-average slip velocity, u
s
, inferred from PEPT
measurements at N
js
, are presented in Fig. 14 at different heights in the vessel for all conditions investigated.

The difference between the slip velocity profiles of the two particle sizes shown in Fig. 14 can be measured by
the root mean square value, thus:

Page 13 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
14

=
|
|

\
|
=
c
n
i
tip
s s
c
rms
u
u u
n
1
2
2 1
1
(19)

where, as before, subscripts 1 and 2 refer to the 1 mm and 3 mm solids, respectively.

At N
js
and above it, the slip velocity varies considerably from point to point throughout the vessel. For example
at 20 wt% solids, the slip velocity for both particle sizes is within the range 0.01-0.42 ms
-1
. The minimum
values tend to occur near the wall within a layer of fluid approximately one-baffle thick, whilst the maximum
values tend to occur below the lower edge of the impeller, in the vicinity of the position (0.70R, 0.20H)
corresponding approximately to the bottom of the circulation loop. These values are generally very different
from the measured terminal settling velocities in stagnant fluid, i.e., u

= 0.18 ms
-1
and 0.36 ms
-1
for the 1 mm
and 3 mm particles, respectively. Even the global volume-average values of u
s
(0.14 ms
-1
; 0.18 ms
-1
) are
considerably different from the u

values.

At N
js
and at the lower solid concentrations, the larger particles with more inertia display considerably more slip
than the smaller ones. This effect, however, reduces considerably at higher solid concentrations, as shown in the
rms
plot in Fig. 15, because at high concentrations particle-particle interactions become important and
differences in velocity reduce through particle collisions. The effect also reduces at speeds above N
js
; again,
because inter-particle interactions also gain importance at higher speeds.

4.6 Verification of mass balance and mass continuity
The pointwise measurements obtained from PEPT and CFD enabled the solids mass balance throughout the
vessel to be accurately verified, i.e. the measured local values of the two solid volume concentrations balance
the experimental parameter C
j
= V
j

/

V
T
, so that:
j
n
i
i
j
c
C c
n
c

=1
) (
1
(20)
where V
j
is the volume of component j in the vessel and V
T
is the total suspension volume. The identity symbol
appears in Eq. (20) because the local concentration is obtained from Eq. 12 through the measurement of
local values of the occupancy O
j
whose average value is by definition identically equal to 1.
Another important tool for checking the accuracy and reliability of flow data is mass continuity. The net mass
flux through a volume bounded by a closed surface S should be zero, thus:
0
) (

S
j
S u
(21)
Calculations can be made by considering a closed cylindrical surface S with the same vertical axis, base and
diameter as the tank but of a shorter height. Because the vector u is zero over the external surface of the vessel
or is parallel to it, the term uS is zero everywhere except in the horizontal plane across the tank, so that S can
be reduced to such a plane, S
h
, and Eq. (21) is reduced to a zero average of axial velocities across the horizontal
Page 14 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
15

plane considered. Alternatively, a similar situation can be envisaged where S is a closed cylindrical envelope
with the same vertical axis and height as the tank but with a smaller diameter; Eq. (21) then becomes a zero
average of radial velocities across the lateral surface, S
r
, of the considered cylindrical volume. These two types
of surfaces S
h
and S
r
were numerically introduced in the vessel and the continuity test was applied to all the sets
of data collected in this study. The results are summarised in Fig. 16, showing a very good verification of the
mass continuity as the velocity average across the given surface in each case is close to zero, generally less than
0.03u
tip
for PEPT and even less for CFD.

5. Conclusions
The three-component flow fields in mechanically agitated dense binary suspensions have been successfully
modelled by CFD. The local velocity components of all the mixture components agree very well with PEPT
measurements throughout the stirred vessel at N
js
and at speeds above it. This has also led to accurate
predictions of the flow numbers. Furthermore, the Reynolds number similarity was found to hold for all
components of the two-phase flow under all conditions investigated.

The spatial distribution of the solid phase is well predicted except close to the base of the vessel and around the
impeller where it is largely overestimated. However, the predictions improve significantly at higher solid
concentrations, as the performance of the Gidaspow model improves at higher concentrations. However, some
improvement might also be due to the Eulerian assumption of a continuum becoming slightly more realistic.
Generally, the smaller particles are better distributed than the larger ones except when N >> N
js
. The uniformity
of suspension improves sharply as N increases above N
js
, but a plateau is reached at high speeds and it becomes
difficult to reach 100% homogeneity, as it is hard to distribute particles below the impeller and near the free
surface. In addition, uniformity increases as a function of mean solid concentration when suspensions are
considered at the same hydrodynamic mixing regime, i.e. at their N
js
speed or the same multiple of their
respective N
js
value.

The local slip velocity varies widely within the vessel and is generally very different from the particle terminal
settling velocity. At N
js
and at lower solid concentrations, the larger particles with more inertia display
considerably more inter-phase slip than the smaller ones. This effect, however, reduces considerably with an
increase in solid concentration and impeller speed as particle-particle interactions become important.
Acknowledgement
Li Lius PhD research was funded by an Elite scholarship from the University of Birmingham. We are grateful
to Prof. D.J. Parker, Dr X. Fang and J.F. Gargiuli of the Positron Imaging Centre at The University of
Birmingham for their assistance with the PEPT experiments.

Supporting Information Available
Table 2 contains a large amount of data on flow numbers obtained from PEPT measurements and CFD
predictions for all conditions investigated, and has been placed in SUPPORTING INFORMATION in accord
with the publication policy of Industrial & Engineering Chemistry Research. This information is available free
of charge via the Internet at http://pubs.acs.org/.
Page 15 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
16


Nomenclature
c local volume concentration of solids -
C

mean volume concentration of solids -
d
p
particle diameter m
D impeller diameter m
Fl multiphase-phase flow number (Q/ND
3
) -
Fl
(L)
liquid flow number (Q
(L)
/ND
3
) -
Fl
(S1)
flow number (Q
(S1)
/ND
3
) for 1mm solids -
Fl
(S2)
flow number (Q
(S2)
/ND
3
) for 3 mm solids -
H height of suspension m
N impeller rotational speed s
-1

n
c
number of PEPT or CFD grid cells -
N
js
minimum speed for particle suspension s
-1

Q impeller pumping rate m
3
s
-1

r radial distance m
R vessel radius m
Re
imp
impeller Reynolds number (ND
2
/) -
S surface area m
2

S1, S2 indices denoting 1 mm and 3 mm particles, respectively
S
h
S
r
horizontal planar and vertical cylindrical surface areas m
2

T vessel diameter m
u
s
slip velocity m s
-1

u

terminal settling velocity m s


-1

u
tip
impeller tip speed m s
-1

u
r
u
z
u

cylindrical velocity components m s


-1

X mean mass concentration of solids -
z vertical distance m

Greek letters
Page 16 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
17

kinematic liquid viscosity m
2
s
-1

uniformity index -
standard deviation of normalised c -

azimuthal coordinate rad
rms
root mean square

Literature Cited
(1) Unadkat, H.; Rielly, C. D.; Hargrave, G. K.; Nagy Z. K. Application of fluorescent PIV and digital
image analysis to measure turbulence properties of solid-liquid stirred suspensions. Chem. Eng. Res.
Des. 2009, 87, 573-586.
(2) Tamburini, A.; Cipollina, A.; Micale, G.; Ciofalo, M.; Brucato, A. Dense solid-liquid off-bottom
suspension dynamics: Simulation and experiment. Chem. Eng. Res. Des. 2009, 87, 587-597.
(3) Guha, D.; Ramachandran, P.A.; Dudukovic, M.P. Flow field of suspended solids in a stirred tank
reactor by Lagrangian tracking. Chem. Eng. Sci. 2007, 62, 6143-6154.
(4) Kagoshima, M.; Mann, R. Development of a networks-of-zones fluid mixing model for an unbaffled
stirred vessel used for precipitation. Chem. Eng. Sci. 2006, 61, 2852-2863.
(5) Ducci, A.; Yianneskis, M. Turbulence kinetic energy transport processes in the impeller stream of
stirred vessels. Chem. Eng. Sci. 2006, 61, 2780-2790.
(6) Chung, K. H. K.; Barigou, M.; Simmons, M. J. H. Reconstruction of 3-D flow field inside miniature
stirred vessels using a 2-D PIV technique. Chem. Eng. Res. Des. 2007, 85, 560-567.
(7) Chung, K. H. K.; Simmons, M. J. H.; Barigou, M. Angle-resolved Particle Image Velocimetry
measurements of flow and turbulence fields in small-scale stirred vessels of different mixer
configurations. Ind. Eng. Chem. Res. 2009, 48, 1008-1018.
(8) Brunazzi, E.; Galletti, C.; Paglianti, A.; Pintus, S. An impedance probe for the measurements of flow
characteristics and mixing properties in stirred slurry reactors. Chem. Eng. Res. Des. 2004, 82, 1250-
1257.
(9) Barresi, A.; Baldi, G.; Solid dispersion in an agitated vessel. Chem. Eng. Sci. 1987, 42, 2949-2956.
(10) Wang, M.; Dorward, A.; Vlaev, D.; Mann, R. Measurements of gas-liquid mixing in a stirred vessel
using electrical resistance tomography (ERT). Chem. Eng. J. 2007, 77, 93-98.
(11) Rammohan, A. R.; Kemoun, A.; Al-Dahhan, M. H.; Dudukovic, M. P. A Lagrangian description of
flows in stirred tanks via computer-automated radioactive particle tracking (CARPT). Chem. Eng. Sci.
2001, 56, 2629-2639.
(12) Villermaux, J. Trajectory Length Distribution (TLD), a novel concept to characterize mixing in flow
systems. Chem. Eng. Sci. 1996, 51, 1939-1946.
(13) Wittmer, S.; Falk, L.; Pitiot, P.; Vivier, H. Characterization of stirred vessel hydrodynamics by three
dimensional trajectography. Can. J. Chem. Eng. 1998, 76, 600-610.
(14) Barigou, M. Particle tracking in opaque mixing systems: An overview of the capabilities of PET and
PEPT. Chem. Eng. Res. Des. 2004, 82, 1258-1267.
Page 17 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
18

(15) Doucet, J.; Bertrand, F.; Chaouki, J. A measure of mixing from Lagrangian tracking and its application
to granular and fluid flow systems. Chem. Eng. Res. Des. 2008, 86, 1313-1321.
(16) Guida, A.; Fan, X.; Parker, D. J.; Nienow, A. W.; Barigou, M. Positron emission particle tracking in a
mechanically agitated solid-liquid suspension of coarse particles. Trans IChemE, Part A, Chem. Eng.
Res. Des. 2009, 87 (4), 421-429.
(17) Barigou, M.; Chiti, F.; Pianko-Oprych, P.; Guida, A.; Adams, L.; Fan, X.; Parker, D. J.; Nienow, A. W.
Using Positron Emission Particle Tracking (PEPT) to study mixing in stirred vessels: validation and
tackling unsolved problems in opaque systems. J. Chem. Eng. Jpn. 2009, 42, 829-846.
(18) Guida, A.; Nienow, A. W.; Barigou, M. PEPT measurements of solid-liquid flow field and spatial
phase distribution in concentrated monodisperse stirred suspensions. Chem. Eng. Sci. 2010, 65 (6)
1905-1914.
(19) Guida, A.; Nienow, A. W.; Barigou, M. Mixing of dense binary suspensions: multi-component
hydrodynamics and spatial phase distribution by PEPT. AIChE J. 2011, 57 (9), 2302-2315.
(20) Guida, A.; Barigou, M. Lagrangian tools for the analysis of mixing in single- and multi-phase flow
systems. AIChE J. 2012, 58 (1), 31-45.
(21) Oshinowo, L.M.; Bakker, A. CFD modeling of solids suspensions in stirred tanks. Symposium on
Computational Modelling of Metals. Minerals and Materials, TMS Annual Meeting, Seattle, WA,
2002, 17-21.
(22) Fan, L.; Mao, Z. Sh.; Wang, Y. D. Numerical simulation of turbulent solid-liquid two-phase flow and
orientation of slender particles in a stirred tank. Chem. Eng. Sci. 2009, 64 (2) 322-333.
(23) Guha, D.; Ramachandran, P.A.; Dudukovic, M.P.; Derksen, J.J. Evaluation of large eddy simulation
and Euler-Euler CFD models for solids flow dynamics in a stirred tank reactor. AIChE J. 2008, 54 (3),
766-778.
(24) Montante, G.; Micale, G.; Magelli, F.; Brucato, A. Experiments and CFD predictions of solid particle
distribution in a vessel agitated with four pitched blade turbine. Chem. Eng. Res. Des. 2001, 79(8),
1005-1010.
(25) Ochieng, A. Lewis, A. E. Nickel solids concentration distribution in a stirred tank, Minerals
Engineering. 2006, 19 (2), 180-189.
(26) Ochieng, A.; Onyango, M. S. Drag models, solids concentration and velocity distribution in a stirred
tank. Power Techonology. 2008, 181(1), 1-8.
(27) Montante, G.; Magelli, F. Modelling of solids distribution in stirred tanks: analysis of simulation
strategies and comparison with experimental data. Int. J. Comp. Fluid Dyn. 2005, 19 (3), 253-262.
(28) Sardeshpande, M.V.; Juvekar, V.A.; Ranade, V.V. Solid suspension in stirred tanks: UVP
measurements and CFD simulations. Can. J. Chem. Eng. 2011, 89 (5), 1112-1121.
(29) Flethcer, D.F.; Brown, G.J. Numerical simulation of solid suspension via mechanical agitation: effect
of the modeling approach, turbulence model and hindered settling drag law, Int. J. Comut Fluid Dyn,
2009, 23(2), 173-187.
(30) Pianko-Oprych, P.; Nienow, A. W.; Barigou, M. Positron emission particle tracking (PEPT) compared
to particle image velocimetry (PIV) for studying the flow generated by a pitched-blade turbine in single
phase and multi-phase systems. Chem. Eng. Sci. 2009, 64 (23), 49554968.
Page 18 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
19

(31) Fangary, Y. S.; Barigou, M.; Seville, J. P. K.; Parker, D. J. Fluid trajectories in a stirred vessel of non-
Newtonian liquid using Positron Emission Particle Tracking. Chem. Eng. Sci. 2000, 55 (24), 5969-
5979.
(32) Fangary, Y. S.; Barigou, M.; Seville, J. P. K.; Parker, D. J. A Lagrangian study of solids suspension in
a stirred vessel by Positron Emission Particle Tracking (PEPT). Chemical Engineering and
Technology. 2002, 25 (5), 521-528.
(33) Zwietering, T. N. Suspending of solid particles in liquid by agitators. Chem. Eng. Sci. 1958, 8 244-253.
(34) Guida, A.; Nienow, A. W.; Barigou, M. Shannon entropy for local and global description of mixing by
Lagrangian particle tracking. Chem. Eng. Sci. 2010, 65 (10), 2865-2883.
(35) Van Wachem, B. G. M.; Almstedt, A. E. Methods for multiphase computational fluid dynamics. Chem.
Eng. J. 2003, 96 (1-3), 8198.
(36) Ljungqvist, M.; Rasmuson, A. Numerical simulation of the two-phase flow in an axially stirred vessel.
Trans IChemE. 2001, 79 (A), 533-546.
(37) Sha, Z.; Oinas, P.; Louhi-Kultanen, M.; Yang, G.; Palosaari, S. Application of CFD simulation to
suspension crystallization-factors affecting size-dependent classification. Powder Technology. 2001,
121 (1), 20-25.
(38) Gidaspow, D. Multiphase flow and fluidization: Continuum and kinetic theory description. Academic
Press: New York, 1994.
(39) Bouillard, J. X.; Lyczkowski, R. W.; Gidaspow, D. Porosity distribution in a fluidized bed with an
immersed obstacle. AIChE J. 1989, 35 (6), 908-923.
(40) Thomas, D. G. Transport characteristics of suspension. Part 8. A note on the viscosity of Newtonian
suspensions of uniform spherical particles. J. Colloid Sci. 1965, 20, 267-277.
(41) Eesa, M.; Barigou, M. Horizontal laminar flow of coarse nearly-neutrally buoyant particles in non-
Newtonian conveying fluids: CFD and PEPT experiments compared. Int. J. Multiphase Flow. 2008, 34
(11), 9971007.
(42) Eesa, M.; Barigou, M. CFD investigation of the pipe transport of coarse solids in laminar power law
fluids. Chem. Eng. Sci. 2009, 64 (2), 322-333.
(43) Deglon, D. A.; Meyer, C. J. CFD modelling of stirred tanks: Numerical considerations. Minerals
Engineering. 2006, 19 (10), 1059-1068.
(44) Coroneo, M.; Montante, G.; Paglianti, A.; Magelli, F. CFD prediction of fluid flow and mixing in
stirred tanks: Numerical issues about the RANS simulations. Computers&Chemical Engineering. 2011,
35, 1959-1968.
(45) Ranade, V. V.; Perrard, M.; Le Sauze, N.; Xuereb, C.; Bertrand, J. Trailing vortices of Rushton turbine:
PIV measurements and CFD simulations with snapshot approach. Trans IChemE, Part A, Chem. Eng.
Res. Des. 2001, 79 (1), 3-12.
(46) Aubin, J.; Fletcher, D. F.; Xuereb, C. Modeling turbulent flow in stirred tanks with CFD: the influence
of the modelling approach, turbulence model and numerical scheme. Experimental Thermal and Fluid
Science. 2004, 28, 431-445.
(47) Van Der Molen, K.; Van Mannen, H. R. E. Laser-Doppler measurements of the turbulent flow in
stirred vessels to establish scaling rules. Chem. Eng. Sci. 1978, 33 (9), 1161-1168.
Page 19 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
20

(48) Venneker, B. C. H.; Derksen, J. J.; Van den Akker, H. E. A. Turbulent flow of shear-thinning liquids in
stirred tanks-the effects of Reynolds number and flow index. Chem. Eng. Res. Des. 2010, 88, 827-843.
(49) Nienow, A.W. The dispersion of solids in liquids, in Mixing of liquids by mechanical agitation, J.J.
Ulbrecht and G.K. Patterson, eds., Gordon & Breach, New York, 1985, 273-307
(50) Gabriele, A.; Nienow, A. W.; Simmons, M. J. H. Use of angle resolved PIV to estimate local specific
energy dissipation rates for up- and down-pumping pitched blade agitators in a stirred tank. Chem. Eng.
Sci. 2009, 64, 126-143.
(51) Hockey, R. M., Nouri, J. M. Turbulent flow in a baffled vessels stirred by a 60 pitched blade impeller.
Chem. Eng. Sci. 1996, 51,4405-4421.
(52) Atiemo-Obeng, V. A.; Penney, W. R.; Armenante, P. SolidLiquid mixing. Handbook of industrial
mixing: Science and practice. Hoboken, NJ; Wiley-Interscience: 2004, 543-584.




Table 1. Range of PEPT measurements and CFD simulations.

X (wt%) C (vol%)

N (rpm) (Re
imp
10
-5
)
N = N
js
N > N
js

Binary
5 2.4 380 (1.51) 570 (2.27) 760 (3.02)
10 4.9 450 (1.79) 675 (2.68) 800 (3.18)
20 10.4 510 (2.03) 765 (3.04) -
40 23.6 610 (2.43) 765 (3.04) -
Mono
d
p
= 1 mm 5.2 2.5 360 (1.43) 450 (1.79) 540 (2.15) 720 (2.86)
20 1.4 490 (1.91) 613 (2.44) - -
d
p
= 3 mm 5.2 2.5 360 (1.43) 450 (1.79) 540 (2.15) 720 (2.86)
10.6 5.2 410 (1.61) 520 (2.07) - 720 (2.86)
20 10.4 490 (1.91) 613 (2.44) - 735 (2.92)












Page 20 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
21



















Figure 1. Computational grid section in the 45 plane between two baffles.






Page 21 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
22












liquid phase d
p
= 1 mm d
p
= 3 mm




Figure 2. Azimuthally-averaged axial distributions of the normalised velocity components of the liquid
and solid components at N
js
: X = 5 wt%; N
js
= 380 rpm.












Page 22 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
23















liquid phase d
p
= 1 mm d
p
= 3 mm




Figure 3. Azimuthally-averaged radial distributions of the normalised velocity components of the liquid
and solid components at N
js
: X = 5 wt%; N
js
= 380 rpm.









Page 23 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
24














liquid phase d
p
= 1 mm d
p
= 3 mm




Figure 4. Azimuthally-averaged axial distributions of the normalised velocity components of the liquid
and solid components at N
js
: X = 10 wt%; N
js
= 450 rpm.










Page 24 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
25













liquid phase d
p
= 1 mm d
p
= 3 mm




Figure 5. Azimuthally-averaged axial distributions of the normalised velocity components of the liquid
and solid components at N
js
: X = 20 wt%; N
js
= 510 rpm.











Page 25 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
26













liquid phase d
p
= 1 mm d
p
= 3 mm




Figure 6. Azimuthally-averaged axial distributions of the normalised velocity components of the liquid
and solid components at N
js
: X = 40 wt%; N
js
= 610 rpm.











Page 26 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
27













liquid phase d
p
= 1 mm d
p
= 3 mm




Figure 7. Azimuthally-averaged axial distributions of the normalised velocity components of the liquid
and solid components at N = 1.8N
js
: X = 10 wt%; N = 800 rpm.











Page 27 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
28













liquid phase d
p
= 1 mm d
p
= 3 mm




Figure 8. Demonstration of Reynolds number similarity Azimuthally-averaged radial distributions of the
normalised velocity components of the three mixture components based on PEPT measurements: X = 5 wt%;
--- X = 10 wt%; --- X = 20 wt%; X = 40 wt%; N = N
js
; Re
imp
= 1.510
5
2.410
5
(see Table 1).










Page 28 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
29
















Figure 9. Azimuthally-averaged axial distributions of normalised solid volume concentration
in suspension at N
js
.






Page 29 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
30















Figure 10. Azimuthally-averaged radial distributions of normalised solid volume concentration
in suspension at N
js
.







Page 30 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
31
















Figure 11. Azimuthally-averaged axial distributions of normalised solid volume concentration
in suspension at N > N
js
.







Page 31 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
32

















Figure 12. Variation of suspension uniformity index as a function of impeller speed for different particle sizes
in binary- and mono-disperse suspensions based on PEPT measurements.








Page 32 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
33
















Figure 13. Comparison of azimuthally-averaged axial and radial distributions of normalised solid volume
concentration based on PEPT measurements in binary- and mono-disperse suspensions at N
js
.












Page 33 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
34

















Figure 14. Azimuthally-averaged radial distributions of normalised time-average slip velocity based on PEPT
measurements at different solid concentrations and agitation speeds.






Page 34 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
35


















Figure 15. Measure of global difference in slip velocity between the two solid components in the vessel as a
function of mean solid concentration based on PEPT measurements.





Page 35 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
36










0
.01
.02
.03
.04
.05
.06
<
u
<
u
<
u
<
u
r rr r
> >> >
S SS S
r rr r
/
u
/
u
/
u
/
u
t tt t
i ii i
p pp p

(
-
)

(
-
)

(
-
)

(
-
)
0 5 10 15 20 25 30 35 40 45
0
.01
.02
.03
.04
.05
Mean solid mass concentration, X (wt%) Mean solid mass concentration, X (wt%) Mean solid mass concentration, X (wt%) Mean solid mass concentration, X (wt%)
<
u
<
u
<
u
<
u
z zz z
> >> >
S SS S
z zz z
/
u
/
u
/
u
/
u
t tt t
i ii i
p pp p

(
-
)

(
-
)

(
-
)

(
-
)


PEPT 1mm
PEPT 3mm
CFD 1mm
CFD 3 mm
S SS S
z zz z
S SS S
r rr r



Figure 16. Normalised radial and axial velocities averaged on surfaces S
r
(of diameter 0.5T) and S
h
(0.3H off
the base), respectively, showing excellent verification of mass continuity by both PEPT and CFD.



Page 36 of 36
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

You might also like