Static Mixer
Static Mixer
Static Mixer
PII: S1004-9541(19)30868-7
DOI: https://doi.org/10.1016/j.cjche.2019.09.004
Reference: CJCHE 1563
Please cite this article as: M.M. Haddadi, S.H. Hosseini, D. Rashtchian, et al., Comparative
analysis of different static mixers performance by CFD technique: An innovative mixer,
Chinese Journal of Chemical Engineering(2019), https://doi.org/10.1016/
j.cjche.2019.09.004
This is a PDF file of an article that has undergone enhancements after acceptance, such
as the addition of a cover page and metadata, and formatting for readability, but it is
not yet the definitive version of record. This version will undergo additional copyediting,
typesetting and review before it is published in its final form, but we are providing this
version to give early visibility of the article. 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.
b
Department of Chemical Engineering, Ilam University, Ilam 69315–516, Iran
c
Department of Chemical Engineering, University of the Basque Country (UPV/EHU), Sarriena s/n, Leioa, Spain.
Abstract: The flow and mixing behavior of two miscible liquids has been studied in an innovative static
of
mixer by using CFD, with Reynolds numbers ranging from 20 to 160. The performance of the new
ro
mixer is compared with those of Kenics, SMX, and Komax static mixers. The pressure drop ratio (Z-
-p
factor), coefficient of variation (CoV) and extensional efficiency (𝛼) features have been used to evaluate
re
power consumption, distributive mixing and dispersive mixing performances, respectively, in all mixers.
The model is firstly validated based on experimental data measured for the pressure drop ratio and the
lP
coefficient of variation. CFD results are consistent with measured data and those obtained by available
na
correlations in the literature. The new mixer shows a superior mixing performance compared to the
other mixers.
ur
Jo
Keywords: CFD, New Static mixer, Coefficient of Variation, Pressure drop ratio.
*
. Corresponding author: E-mail address: [email protected] (S. H. Hosseini)
Journal Pre-proof
1. Introduction
Mixing is one of the most important unit operations in many processes, which has a major role
on the quality of the final product. Improper mixing can result in a lower quality of products, and
therefore the need for more sophisticated separation systems, which in turn increases waste
especially in flows with low Reynolds numbers, as they involve non-homogenous and
of
imperfectly mixed flow. Mixing is carried out as a batch operation in agitated vessels and as a
continuous operation in static mixers [1]. The static mixers are composed of numerous
ro
motionless elements installed in a fixed pipe or channel. The aim of these elements is to split,
-p
rotate and then recombine the input streams. Since there are no moving parts in static mixers, the
re
only power required is the one of the external pump in order to overcome the pressure loss
lP
generated by the mixing elements and fluid movement. Some benefits of static mixers in
comparison with dynamic ones can be listed as follows: small space required for installation, low
na
maintenance costs, low power consumption, absence of moving parts, short residence times of
ur
the fluids, near plug flow behavior of the fluids, efficient heat and mass transfer, good mixing at
low shear rate, and capability for self-cleaning [2-4]. The static mixers are used in many
Jo
processing applications, such as mixing of miscible liquids [5], immiscible liquids [6,7],
dispersion of viscous liquids [8], mass transfer [9], heat transfer [10], and gas-liquid dispersion
[11,12]. These applications are available in many industries, such as polymer processing [13],
biotechnology [14], wastewater treatment [15-18], food processing [19], and petrochemistry
[20,21]. The wide range of applications of static mixers in various industries is evidence of the
2
Journal Pre-proof
The performance of a static mixer can be mainly specified by its mixing efficiency and pressure
drop. First, they are assessed by the pressure drop ratio (pressure drop of the mixer to pressure
drop of the empty pipe). Then mixing efficiency parameters are considered by using the
coefficient of variation, the extensional efficiency, and the residence time distribution. The
mixing occurs, droplets or bubbles are distributed in the mixer. This mechanism differs from
dispersive mixing, which corresponds to size reduction [22]. In order to perform distributive
of
mixing, a component A is mixed in a mainstream with another component B. Then, the
ro
concentration of component A is monitored and mapped as a function of position in a specified
-p
cross-section. Afterwards, the standard deviation of the concentration is calculated and then the
re
coefficient of variation (CoV) is specified. If component A is uniformly distributed in a cross-
section, the CoV tends to zero. The residence time distribution is sometimes used as a mixing
lP
criterion, especially when the systems include a chemical reaction. A narrower distribution leads
na
to better mixing [23]. Moreover, the extensional efficiency is introduced in order to investigate
In the last few decades, computational fluid dynamics (CFD) has become an essential numerical
Jo
tool in the prediction of fluid dynamics, heat transfer and also mass transfer in packed beds,
which are used in various industries [25,36]. Therefore, CFD can be successfully applied to the
Different aspects of the static mixers have been studied both experimentally and numerically. For
instance, Liu et al. [27] and Baumann et al. [28] performed numerical and experimental studies,
respectively, to measure the pressure drop in SMX static mixers. Rauline et al. [22] studied the
velocity field for creeping flow in six static mixers using three-dimensional finite volume CFD
3
Journal Pre-proof
code. They used the numerical results of the extensional efficiency, pressure drop, mean shear
rate, stretching, and coefficient of variation to compare the performances of those mixers. They
proved that the SMX mixer was the most efficient amongst all configurations. Moreover, by
studying the Kenics mixer they showed that this mixer is more efficient when certain space is left
between the mixing elements. Zalc et al. [29,30] investigated the hydrodynamics and mixing
behavior in an SMX mixer at Reynolds number in the range of 10 -4 −100. Their results revealed
that when the injection is made at the center of the tube, the mixing efficiency is higher than for
of
off-center injection. In addition, they showed that the pressure drop ratio does not change for
ro
Reynolds numbers lower than 10. Several CFD simulations of flow behavior through Kenics
-p
mixers are found in the open literature [31-36]. Furthermore, some experimental studies have
re
also been conducted by several researchers on the flow behavior in Kenics static mixers [37,38].
However, there are few numerical studies about the Komax mixer.
lP
Most of the studies about designing novel static mixers or modifying static mixers available at
na
laboratory scale have been limited to creeping flows and laminar flows at higher flow rates, i.e.,
ur
to a low Reynolds number range, in which the influence of internals, such as blade shape,
number of elements and other design factors, have been assessed by numerical tools [34,39-42].
Jo
Regner et al. [34] studied the characterization of two static mixers, namely, Kenics KM and
Lightnin Series 45, by evaluating Z factor, helicity and the rate of striation thinning based on
CFD. The Reynolds number they used was in the 0.1-160 range. Singh et al. [39] analyzed, and
subsequently optimized, the various design parameters of SMX mixers, namely, N p, the number
of parallel cross-bars per element and, Nx, the number of cross-bars over the width of the
channel, by the mapping method (MM) under a low Re number (0.44). They obtained that the
optimum design obeys the rule of Np = (2/3)Nx – 1. To achieve the enhanced mixing
4
Journal Pre-proof
performance of the SMX geometry, Soman and Madhuranthakam41 simulated different SMX
at Re numbers ranging from 30 to 100. They compared the results in terms of velocity field,
pressure field, shear rate and extensional efficiency for dispersive mixing, and the standard
deviation of the concentration for distributive mixing. They found that circular serrations in
SMX mixers reduce the pressure drop by 33–35% of that corresponding to the standard SMX
mixer. Meng et al. [41] compared (using a CFD tool) the mixing performance of Kenics and
of
Lightnin static mixers for a high-viscosity fluid at low Reynolds numbers in the 0.1-100 range
ro
and different aspect ratios, as well as different elements. They proposed a new mixing factor ƞ to
-p
evaluate the micro-mixing performance in the mixers. In another work, Meng et al. [42] studied
re
the mixing performance of a high-viscosity fluid in novel static mixers with multitwisted leaves
In addition to their use for the laminar flow, some numerical works in the open literature deal
na
with novel static mixers or modifications to the existing ones for use in turbulent flow. For
ur
instance, Meng et al. [43] simulated perforations in the standard Kenics mixer by CFD tool and
studied the impact of their size and spacing, as well as the number of segments, on the heat
Jo
transfer under the turbulent condition. They achieved outstanding results using d/W= 0.3 and
s/W= 0.6 in terms of friction loss and heat transfer rate in the modified Kenics mixer. Zhou et al.
[44] proposed a new static mixer of hollow cross disk in order to increase the turbulent reactive
Few studies have been reported in the literature regarding static mixers with complex fluids, such
as non-Newtonian and immiscible ones. In this respect, Jegatheeswaran et al. [45] studied the
intensification of non-Newtonian fluid mixing in chaotic SMX static mixers at low Reynolds
5
Journal Pre-proof
numbers. They also reviewed process intensification by mixing non-Newtonian fluids in static
mixers, and suggested to perform CFD studies to shed light on the impact the geometry and
design parameters of static mixers have on fluid deformation and mass transfer under low
Reynolds numbers for both Newtonian and non-Newtonian fluids [46]. Mihailova et al. [47]
evaluated a Newtonian fluid (glycerol) and a non-Newtonian one (guar gum solution) in a SMX
mixer, and found that their velocity distribution is hardly affected by rheology at the same low
flow rate. Concerning the performance of immiscible fluids in static mixers, few studies have
of
been reported in the literature, and they use Eulerian-Lagrangian and Eulerian-Eulerian
ro
approaches combined with the population balance in order to study particle trajectories and
mixer on the hydrodynamics and mixing behavior of brine 25% (wt) by using CFD in four static
mixers at low Reynolds numbers. In view of the brief review done above, the mixing behavior of
na
the new static mixer is firstly analyzed at low Reynolds numbers due to the high probability of
ur
having fluid inhomogeneity under those conditions. Accordingly, this study covers a wide range
the new simple static mixer inspired by the behavior of dynamic mixers in vessels. Amongst the
different criteria, extensional efficiency, pressure drop ratio and the coefficient of variation have
been chosen to evaluate the performance of the well-known Kenics, SMX and Komax mixers
and of an innovative static mixer. The simulation results are validated by comparing the
predicted results with the existing correlations for the pressure drop ratio in a Kenics mixer and
the experimental data for the coefficient of variation in a mixer with baffle shape elements.
2. Numerical methodology
6
Journal Pre-proof
This study analyses the performance of four static mixers, namely, Kenics, SMX, Komax, and a
new static mixer proposed by the authors. Furthermore, in order to validate the CFD model,
another two other static mixers have been used, namely, a static mixer with staggered
semicircular baffle shape elements [50] and a Kenics static mixer [51] with the same geometrical
dimensions and operating conditions as those used in the mentioned four static mixers. Based on
of
the previous works [50,51], it is assumed that the tube and internal elements of the static mixers
are made of Plexiglass. Figure 1 shows the computational domains of Kenics, SMX, Komax and
ro
the new static mixer. Moreover, Figure 2a illustrates the geometry of the blades in the new
-p
mixer. More information about the static mixer with baffle shape elements can be found
re
elsewhere [50]. In Figure 1, the computational domain comprises a horizontal tube with a
lP
diameter of D and four mixing elements. Two empty tubes of 3D length are added prior and
subsequent to the mixing elements in order to provide developmental flow length and to avoid
na
creating backflow in the simulations. As shown in Figure 2b, the injection tube is a curved 3 mm
ur
diameter tube aligned with the main axis, with the injection point being located at a distance of
15 mm from the tube inlet. The detailed dimensions of the static mixers and the properties of the
Jo
fluid are set out in Tables 1 and 2, respectively. Each element of the Kenics mixer is a twisted
plate that is alternated and oriented so that each leading edge is at 90° of the trailing edge of the
next one. The mixing elements in the SMX mixer comprise inclined bars crossed to one another.
Each element is rotated 90° with respect to the foregoing element. The Komax static mixer is
manufactured from slotted plates with inclined ends. The flow channels with triangular shapes
7
Journal Pre-proof
of
ro
-p
re
lP
na
ur
Jo
Figure 1. Geometries of various static mixers, namely, Kenics mixer, SMX mixer, Komax mixer,
and novel mixer in different views.
The first mixing element of the new mixer consists of two rows of blades at a distance of D/2
from each other with 90° differences in angle. Each row has four blades, which form two crosses
8
Journal Pre-proof
with an angle of 45° with respect to the tube axis at a distance of D/2 from each other. The
second element of the new mixer is the same as the first element with the difference that all
blades are rotated 90° around the pipe axis. Thus, the number of required elements to construct
the new mixer should be at least two. It is interesting to note that the design of the proposed
mixer is inspired by the dynamic mixers in stirred vessels, which include several blades at
of
ro
-p
re
lP
na
ur
Jo
Figure 2. (a) Geometry of the blades in the new static mixer (b) sketch of the pipeline equipped
with mixing elements.
The general assumptions of the model used are as follows: the flow is single-phase, three-
dimensional, steady state, incompressible, isothermal and without any reaction (without mass
production or mass consumption). The continuity, momentum, and species transport equations
9
Journal Pre-proof
are solved by a finite volume method (FVM). The mentioned equations are expressed as follows
[52]:
∇. (𝜌𝐕) = 0 (1)
of
where 𝜇, 𝜌, 𝐠, 𝐕, P, Y𝐴 , DAB stand for the bulk dynamic viscosity, bulk density, gravity
ro
acceleration vector, mean velocity vector, pressure, mass fraction of component A, and
-p
Table 1. Geometrical Parameters of Static Mixers
re
parameter Kenics SMX Komax New mixer
lP
10
Journal Pre-proof
equations. The origin of the z-axis is positioned at the inlet of the tube and its direction is
opposite to the main flow. The convective terms in the governing equations are discretized by a
of
second order upwind scheme, and the SIMPLE algorithm is used for pressure-velocity coupling.
ro
Gradients are essential to compute diffusion terms and velocity derivatives. Moreover, the values
-p
of the scalar at the cell faces should be created. To solve the conservation equations and calculate
re
the gradients, a commercial ANSYS FLUENT 16.0 software is used in order to apply a Least-
Square cell-based discretization method. The convergence limit is set to a scaled residual lower
lP
than 10-5 for all the mentioned equations. The various geometries of the mixers and the
na
corresponding computational grids have been created by ANSYS DESIGN MODELER and
ANSYS MESHING, respectively. The simulations are performed on an Intel Core™ i7-6700HQ
ur
̅D
𝜌V
Re = 𝜇
(4)
̅ stands for the mean velocity and D for the diameter of the tube including the mixing
where V
elements. The boundary condition at the inlet of the mixer is a fully developed velocity profile,
which has been introduced into the software using a user-defined function (UDF). According to
11
Journal Pre-proof
previous modeling of static mixers [5,53,54], the no-slip boundary condition is utilized on the
tube wall and element surfaces. The atmospheric pressure, i.e., zero gage pressure, is utilized at
the outlet of the static mixers. Brine 25% (wt) is injected into the tube co-currently to the main
flow. The tracer (brine) velocity chosen at each Reynolds number (based on the inlet condition of
the main liquid) is that corresponding to a ratio of 0.1 between the tracer flow rate and the main
flow rate.
of
2.4. Grid independence test
ro
Grid independence test is performed to determine the proper grid size, ensuring the numerical
-p
results are independent of the grid system. The maximum values of Equisize skew for all
geometry meshes are lower than 0.79, which is evidence of the good quality of the meshes based
re
on the FLUENT criterion. In addition, the geometries and grids have been created by
lP
DesignModeler, from ANSYS productions. In order to perform the grid independence test, four
static mixers (Kenics, Komax, SMX and the novel one) with a length of 300 mm and a diameter
na
of 25 mm are used with six different grid sizes ranging from 0.5 to 3 mm (0.5 mm is chosen for
ur
the step size) at Re = 160. Tables 3 to 6 show the results of the grid independence test. As
Jo
shown in these tables, both Z-factor (related to pressure drop) and CoV (related to the
concentration at the plate located 1.5D downstream) decrease as the grid size is increased above
1.5 mm. In addition, by changing the grid size from 1 to 0.5 mm, Z-factor and CoV change by
less than 0.7 % and 0.5 %, respectively. Tables 3 to 6 summarize the key characteristics of the
grids, and they clearly show that CFD simulation time is highly dependent on the grid size.
Given the slight difference between 1 and 0.5 mm grid sizes, medium grid size (1 mm) strikes
the balance between accuracy and computational time, and has been used for further simulations.
12
Journal Pre-proof
Table 3. Grid size, CFD simulation time, number of elements, estimated Z and CoV for the Kenics mixer at Re=160.
of
elements
ro
time (min)
Simulated CoV
Simulated Z
0.185
10.23
0.187
10.22 -p
0.15
10.27
0.07
10.14
0.04
9.95
0.02
9.85
re
lP
Table 4. Grid size, CFD simulation time, number of elements, estimated Z and CoV for the Komax mixer at
Re=160.
na
elements
time (min)
Table 5. Grid size, CFD simulation time, number of elements, estimated Z and CoV for the SMX mixer at Re=160.
13
Journal Pre-proof
Table 6. Grid size, CFD simulation time, Number of elements, estimated Z and CoV for the new mixer at Re=160.
of
elements
ro
time (min)
Simulated CoV
Simulated Z
0.001997
63.2
0.002
63.5 66-p
0.0017 0.0015
64.2
0.0012
62
0.001
60.5
re
lP
The accuracy of the numerical model is an essential aspect to be quantified prior to use it for
predicting new situations. To assess the validity of the CFD results, the simulated pressure drop
ur
ratio and the coefficient of variation are compared with the experimental data and available
Jo
correlations. The Z-factor is defined as the ratio of the static mixer pressure drop, ∆P, to the
∆P
Z = ∆P (5)
0
In this study, the pressure drop is calculated as the absolute difference between the area-weighted
average pressure at the planes located at 1.5D upstream and downstream the mixing elements.
14
Journal Pre-proof
This is an important parameter for static mixers and provides a measure of the energy required
∑N
i=1(Yi −Ymean )
2 1
CoV = √ (6)
N−1 Ymean
of
where Yi is the local mass fraction of the tracer at the ith point, N is the number of evaluation
ro
points and Ymean is the mean mass fraction of the tracer on the cross-section of the tube.
Therefore, values of the CoV are calculated in various points of the tube cross-section following
-p
previous works [18,50]. Figure 3(a, b) compares the values computed for Z-factor in the range
re
Re = 20 − 160 and for CoV in the range Re = 70 − 480 with the data measured in the Kenics
lP
mixer [52] and in a static mixer with baffle shape elements [50], respectively. Figure 3a shows
that an increase in the Reynolds number leads an increase in pressure drop, which in turn leads to
na
an increase in Z-factor.
ur
Jo
15
Journal Pre-proof
of
ro
-p
Figure 3. (a) Experimental and computed pressure drop ratios for various Reynolds numbers (b)
Experimental and computed CoV values for various Reynolds numbers.
re
1 |Simulationi −Experimentali |
MRE = N ∑N
lP
i=1 |Experimental i |
(7)
The mean relative error (MRE) between the data measured by Sir and Lecjaks [51] and the CFD
na
results is approximately 2.3%. Accordingly, a quite good quantitative agreement between the
ur
simulated data and the experimental ones by Sir and Lecjaks [51] is found for Z-factor. Figure 3b
compares the computed values of CoV with the corresponding experimental data provided by
Jo
Al-Atabi [50] in a static mixer with baffle shape elements for Re = 70 − 480. The MRE
between the experimental data (Al-Atabi, 2011) and the simulated results is of around 6%.
Therefore, a good agreement between the results is also attained for CoV. Overall, Figure 3(a, b)
shows that the current model is able to predict the hydrodynamics and mass transfer of brine into
water flowing through the various static mixers at different Reynolds numbers with a reasonable
accuracy. As a result, the current CFD model is suitable for studying the other static mixers.
Figure 4 compares the CFD results with those for well-known available correlations developed
by Wilkinson and Cliff [55], Cybulski and Werner [19], Sir and Lecjaks [51], Shah and Kale
[56], and Grace [57] for a Kenics mixer including four elements. These empirical expressions
that are commonly used for estimating Z-factor in Kenics static mixers are listed in Table 7.
There is a considerable discrepancy in the results predicted by the different correlations for Z-
factor. This discrepancy is explained by the differences in the dimensions of the static mixers,
such as diameter and length, and because the impact of some geometrical features has been
of
ignored in the correlations. Unlike the variability predicted by the existing correlations for the
ro
Kenics static mixer, the numerical results are consistent. In the mentioned range of Reynolds
-p
numbers, the computed Z-factors have 3.6% and 8% MRE compared to the correlations by
re
Cybulski and Werner [19], and Sir and Lecjaks [51], respectively.
lP
na
ur
Jo
17
Journal Pre-proof
of
ro
-p
re
lP
Figure 4. Comparison of CFD results with those calculated using the correlations available for
the Kenics mixer at various Reynolds numbers.
na
One of the most important sources responsible for numerical errors in CFD is numerical
diffusion. This type of error stems from the discretization of convective terms in the governing
18
Journal Pre-proof
equations. To investigate the numerical diffusion and select the appropriate discretization scheme
for the convective terms, five discretization schemes (First order upwind, Second order upwind,
QUICK, MUSCL and Power law) have been compared with each other. The results of this
comparison are shown in Figure 5. This figure clearly shows that except the first order upwind
and power law schemes, the other discretization schemes lead to similar results. Therefore, the
use of QUICK, MUSCL and second order upwind schemes do not lead to significant error in the
computation. The percentages have been calculated for the numerical differences between the
of
accurate scheme of second order upwind and the other schemes. The minimum and maximum
ro
differences for mass transfer are obtained by the QUICK scheme with a MRE of 0.37 % and by
-p
the first order upwind with a MRE of 78 %, respectively. The minimum and maximum
re
differences for pressure are obtained by the QUICK scheme with a MRE of 0.64 % and by the
first order upwind with a MRE of 45%, respectively. In addition, the second order upwind
lP
scheme was specifically used in many studies to simulate flow in static mixers and other systems
na
[32, 58, 59]. Accordingly, the second order upwind discretization scheme has been used in this
19
Journal Pre-proof
of
Figure 5. Radial distribution of (a) total pressure, (b) mass fraction of the brine at the outlet of
ro
Kenics static mixer with four mixing elements.
four static mixers has been evaluated and the results are discussed for various conditions.
na
The contour plots of cross-sectional velocity profiles for different static mixers are shown in
Figure 6(a-d) at different locations from the inlet of the first mixing element to the end of each
Jo
mixing element at Re = 60. This figure clearly shows that the maximum values of the fluid
velocity are quite different in the mixers due to their different structure. Figure 6a shows the
contour of fluid velocity in the Kenics mixer. As observed, four velocity zones are formed at the
end of the first element. As the flow passes through the mixer, these zones combine together and
create new zones with different velocities. As expected from the mass conservation, the velocity
is proportional to the Reynolds number and is the same at the end of the first and third mixing
20
Journal Pre-proof
elements. In the fourth element, the four velocity zones combine together again and two velocity
of
ro
-p
re
lP
na
ur
Jo
Figure 6. Cross sectional profiles of velocity at Re=60 for: (a) Kenics mixer (b) Komax mixer (c)
Proposed mixer (d) SMX mixer.
Figure 6b shows the velocity distribution in the Komax mixer. When the fluid flow enters the
first mixing element, it splits into two streams. As the cross-sectional area of each stream
decreases, the stream velocity increases. It is obvious from Figure 6b that a high velocity zone
near the wall mixer is created in each stream. This is due to the rotation of the fluid, which
21
Journal Pre-proof
causes a centrifugal force and creates high speed zones at a distance far from the center of
rotation. This phenomenon is also observed in the Kenics mixer. The velocity profiles of the new
mixer are shown in Figure 6c. Similarly to the Komax mixer, there are two zones with high
velocity in this mixer. It is remarkable that these zones are created when the fluid passes through
the first mixing element, which is not the case in the Komax mixer where they are created after
the second mixing element. This means that the new mixer can create more disturbances
compared to the Komax one. In addition, the velocity contours of the SMX mixer are shown in
of
Figure 6d. As can be seen, the cross-bars divide the input fluid into several streams with different
ro
velocities. Furthermore, the high velocity zone in the SMX mixer is only in a region after the
Figure 7a shows the values of Z-factor computed for different static mixers with the same
na
dimensions and four mixing elements. As expected, Z-factor increases by increasing the
ur
Reynolds number. It is clearly seen that there is a great difference in Z-factor for the different
Jo
static mixers. The Z-factor in the SMX mixer is higher than in the other mixers. This is due to
the complex geometry and the small space between the cross-bars in the SMX mixer. Moreover,
Figure 7a shows that the Z-factor for the new mixer is about 45% lower than for the SMX mixer
at all Reynolds numbers, whereas this factor is higher than those in the Kenics and Komax
mixers. In addition, the Z-factor of the new mixer has a lower slope than for the SMX mixer.
Thus, it is concluded that even for Reynolds numbers above the range studied, the Z-factor of the
new mixer is lower than that of the SMX mixer, with the trends of Z-factors being approximately
parallel in these mixers. Nevertheless, the lowest computed Z-factor corresponds to the Kenics
22
Journal Pre-proof
mixer. In order to investigate more accurately pressure drop in the static mixers, their Z-factor
for attaining a value of CoV=0.05 have been plotted in Figure 7b. As shown in Figure 7b, when
the Reynolds number is greater than 40, the new mixer leads to a lower pressure drop compared
to the other mixers to achieve CoV=0.05. Moreover, high mixing qualities in the Kenics mixer
require high pressure drops. This difference in the pressure drop produced by the mixers is due to
the difference in the number of mixing elements required to reach a given degree of mixing.
Another point to be noted in Figure 7b is the decrease in Z-factor when the Reynolds number
of
(flow rate) is increased. In fact, an increase in flow rate increases the rate of mixing, and
ro
therefore less mixing elements are required and so pressure drop decreases.
-p
re
lP
na
ur
Jo
Figure 7. (a) Comparison of Z-factor in the different static mixers with four mixing elements, (b)
Z-factors in the static mixers for different Reynolds numbers when the target is CoV=0.05.
23
Journal Pre-proof
Figure 8a shows the relationship between the Reynolds number and the coefficient of variation
for the different static mixers. From this figure, it is clear that an increase in Reynolds number
leads to a reduction in CoV. At all ranges of Reynolds number, the Kenics mixer has the highest
CoV value. Moreover, for Re < 40 the new static mixer provides a higher CoV in the tube cross-
section than the SMX mixer, and therefore has a lower mixing efficiency. Nevertheless, for
Re > 40, the new mixer shows a better performance in the distributive mixing compared to the
other static mixers. Figure 8b shows the relationship between the coefficient of variation and the
of
length of the mixer at a constant Re number of 60. By increasing the mixer length, the CoV value
ro
decreases gradually. For L/D > 9 in the SMX and new mixer, CoV reaches almost a constant
-p
value. In addition, the new mixer shows the best performance in the distributive mixing
re
compared to the other mixers for all ranges of L/D. Overall, the new static mixer shows an
attractive mixing performance because pressure drop is considerably lower (about 45%) than for
lP
the SMX mixer. Therefore, to achieve a desired mixing efficiency, the new static mixer leads to
na
24
Journal Pre-proof
of
ro
-p
Figure 8. Comparison of CoV values in the different static mixers: (a) Influence of the Reynolds.
number on the coefficient of variation (b) Effect of L/D on the coefficient of variation at Re=60.
re
lP
One of the most important parameters in the evaluation of dispersive mixing is extensional
|γ|
𝛼 = |γ|+|ω| (8)
Jo
where |ω| and |γ| are the norms of the rate of vorticity and deformation tensors, respectively,
1
|γ| = [∇V + ∇V T ] (9)
2
1
|ω| = [∇V − ∇V T ] (10)
2
25
Journal Pre-proof
where ∇V is the velocity gradient. Generally, a higher value of 𝛼 means a higher dispersive
The general trend in a static mixer is that a fluid with a high extensional efficiency has both high
velocity gradient and high deformation tensor. According to the correlation shown below (Eq.
(11)) proposed by Das et al. [60], as the velocity gradient, and therefore the extensional
efficiency, is higher the static mixer may produce finer droplets in multi-phase systems.
of
2𝜎
𝑑𝑚𝑎𝑥 = 𝐶𝑎𝑐𝑟𝑖𝑡 𝛾̇ 𝜇 (11)
𝑐
ro
where 𝑑𝑚𝑎𝑥 is the maximum equilibrium droplet diameter, 𝐶𝑎𝑐𝑟𝑖𝑡 is the critical capillary
-p
number, 𝜎 is the surface tension between the dispersed and continuous phases, 𝜇𝑐 is the
continuous phase molecular viscosity and 𝛾̇ is the strain rate, which depends on the velocity
re
gradient.
lP
na
ur
Jo
26
Journal Pre-proof
of
ro
-p
re
lP
na
ur
Jo
Figure 9. Extensional efficiency of the (a) Kenics mixer (b) Komax mixer (c) Proposed mixer (d)
SMX mixer.
Figure 9 shows the extensional efficiency over a line with coordinates (x = D/6), (y = D/6),
(−0.225 < z < −0.075 m) for various static mixers. The horizontal axis represents the number
of elements. As observed in Figure 9a, the parts between the elements in the Kenics static mixer
cover greater areas for higher values of 𝛼, which means that the parts between two mixing
element groups have a significant effect on the dispersive mixing. For the Komax mixer (Figure
9b), the maximum value of 𝛼 is observed at element 0 and element 2, which corresponds to the
27
Journal Pre-proof
inlet of the first and third mixing elements, respectively. Three smaller peaks are observed at the
mixing elements 1, 3 and 4, which correspond to the inlet of the second and the outlet of the third
and fourth mixing elements, respectively. The wide ranges of extensional efficiency observed in
Figures 9c and 9d reveal that more complex flow patterns are predicted for the SMX and the new
mixer than those for the Kenics and Komax mixers. From the average extensional efficiency
values shown in Figure 9, it can be deduced that the new static mixer is the most efficient one
concerning dispersive mixing, with the Komax one being second, followed by the SMX and the
of
Kenics mixers. Moreover, the two different types of static mixers (the new one and the Komax
ro
mixer) have almost the same values of average extensional efficiency, but different CoV values.
-p
This means that the two mixers that have the same dispersive mixing, but not necessarily lead to
re
the same distributive mixing. The average value is not the only way to analyze this criterion.
Thus, a high extensional efficiency is required to deform a single-phase fluid or break a drop in a
lP
two-phase fluid.
na
The cumulative volumetric distributions of 𝛼 in Kenics, Komax, SMX and in the new mixer are
ur
shown in Figure 10. These percentages are the volume percentages of the mixers in which the
extensional efficiency is greater than the value in the abscissa. Thus, the extensional efficiency is
Jo
higher than 0.5 for 36% of the Kenics volume, 68% of the Komax volume, 79% of the SMX
volume, and 92% of the new mixer volume. For any value, the Kenics mixer is less efficient than
the Komax and the new mixer, since for any value of 𝛼 the cumulative volumetric percentage is
lower. On the other hand, the SMX mixer is more efficient than the Kenics mixer only if 𝛼 is
below 0.56. Moreover, it can be seen that the new mixer is more efficient than the other static
mixers, except in the range from 0.63 to 0.75, in which its efficiency is slightly lower than that of
28
Journal Pre-proof
of
ro
-p
Figure 10. Cumulative volumetric percentage in the different static mixers.
re
lP
The CFD results of brine concentration distribution at different cross-sections of the static mixers
na
have been evaluated before and after the mixing elements. Figure 11 shows a sample of the
ur
and brine) enter the first element and split into two (Kenics and Komax) or more (SMX and new
mixer) streams along the surfaces of the elements, which promote mixing compared to the flow
at different sections of an empty tube. By analyzing the brine distribution at various cross
sections, it is found that after the third element the homogenization reaches a satisfactory level
(CoV < 0.05) for the SMX and the new mixer, whereas for the Kenics and Komax mixers the
brine was not uniformly distributed across the mixers subsequent to the fourth element. Overall,
the differences in the mixing intensity of the brine shown in Figure 11 are evidence that the
of
ro
-p
re
lP
na
ur
Jo
Figure 11. Mass fraction distribution of brine at different cross sections for different static mixers
at Re=60: (a) Kenics mixer (b) Komax mixer (c) Proposed mixer (d) SMX mixer.
30
Journal Pre-proof
of
ro
-p
re
lP
na
ur
Jo
Figure 12. Mass fraction distribution of brine on a diametrically vertical plane in the static mixers
at Re=60: (a) Kenics mixer (b) Komax mixer (c) Proposed mixer (d) SMX mixer.
Figure 12 shows the contour plots of the mass fraction of the brine at Re = 60 along the static
mixers. As observed in Figure 12, two streams from two different boundaries (the inlet of the
main pipe and the injection tube) enter the first element, and the mixing elements change the
direction of the flow. This issue leads to full mixing of the fluids at the end of the SMX and the
new mixer.
31
Journal Pre-proof
4. Conclusions
A three-dimensional CFD simulation was performed for different static mixers, namely, Kenics,
SMX, Komax and a novel static mixer, for Reynolds numbers ranging from 20 to 160. There are
a number of correlations in the open literature to predict the pressure drop ratio for well-known
static mixers, such as the Kenics one. These correlations have been mainly developed for laminar
and turbulent flow regimes. However, most of these correlations do not consider the geometric
of
parameters, such as the blade thickness and the number of elements, and they are therefore under
great uncertainty. To validate the numerical model, experimental data available for a static mixer
ro
with a baffle shape element (Al-Atabi [50]) and a correlation proposed for the Kenics mixer (Sir
-p
and Lecjaks [51]) have been taken as benchmarks. The pressure drop ratio and the coefficient of
re
variation predicted by the CFD model have been compared with experimental data reported in
lP
the open literature, and a good agreement (2.3% and 6% for Z-factor and CoV, respectively) has
been found. Moreover, some available correlations for Z-factor have been evaluated for a Kenics
na
mixer and the best fit was obtained for the one proposed by Cybulski and Werner [19] for
ur
Re = 20 − 160. After the validation of the CFD model, the computational method was extended
to the other static mixers in order to examine the effect operating and geometric parameters have
Jo
on the hydrodynamics and mixing performances. The velocity field, pressure drop ratio,
determined at various conditions in different mixers. The proposed static mixer has a much better
performance than the other static mixers. Finally, the following remarks may be drawn from this
study:
1) The Z-factor generated in the new static mixer is about 45% lower than that
2) The mixing efficiency of the new mixer is higher than those of the other static mixers.
Actually, this mixer generates a lower coefficient of variation and higher extensional
3) Two mixers with the same dispersive mixing do not necessarily lead to the same
distributive mixing.
4) The new static mixer is more efficient than the other mixers in deforming fluid (single-
phase) and breaking drops (two-phase) because it creates high extensional efficiency.
of
5) The methodology described here can be applied to simulate and design other mixers in
ro
order to evaluate their mixing performances.
-p
Finally, given that the information about use of static mixers with complex fluids is scarce, the
re
mixing performance of this new mixer will be assessed in a future study involving complex
lP
Nomenclature
ur
33
Journal Pre-proof
of
L: length of the mixer, (mm)
ro
N: number of evaluation points, ‒
𝐕: velocity, (m/s)
̅:
V mean velocity, (m/s)
34
Journal Pre-proof
Greek letters
of
α: extensional efficiency
ro
γ: deformation tensor, (1/s)
References
[1] K.J. Myers, A. Bakker, D. Ryan, Avoid Agitation by Selecting Static Mixers, Chem. Eng.
[2] T. Bor, The Static Mixer as a Chemical Reactor, Br. Chem. Eng. 16 (1971) 610−612.
[3] J.R. Baker, Motionless mixers stir up new uses, Chem. Eng. Progress. 87 (1991) 32−38.
35
Journal Pre-proof
[4] R.K. Thakur, Ch. Vial, D.P. Nigam, E.B. Nauman, G. Djelveh, Static Mixers in the Process
[5] G. Montante, M. Coroneo, A. Paglianti, Blending of miscible liquids with different densities
[6] F. Theron, N. Le Sauze, Comparison between three static mixers for emulsification in
of
[7] P. Pianko-Oprych, Z. Jaworski, CFD modelling of two-phase liquid-liquid flow in a SMX
ro
static mixer, Pol. J. Chem. Technol. 11 (2009) 41−49.
-p
[8] P.D. Berkman, R.V. Calabrese, Dispersion of Viscous Liquids by Turbulent Flow in a Static
re
Mixer, AIChE J. 34 (1988) 602−609.
lP
[9] C.U. Ugwu, J.C. Ogbonna, H. Tanaka, Improvement of mass transfer characteristics and
na
[10] R. Rakoczy, S. Masiuk, M. Kordas, P. Gradzik, The effects of power characteristics on the
Jo
heat transfer process in various types of motionless mixing devices, Chem. Eng. Process. 50
(2011) 959−969.
analysis of dispersive mixing by a helical static mixer in upward co-current gas–liquid flow,
liquid flow in a helical static mixer, Chem. Eng. Sci. 137 (2015) 476−486.
36
Journal Pre-proof
[14] E.L. Paul, Design and Scaleup of an Anchorage-dependent Mammalian Cell Bioreactor,
[15] Z. Qing-shi, Ozone Disinfection of Sewage in A Static Mixer Contacting Reactor System on
[16] N. Martin, C. Galey, Use of Static Mixer for Oxidation and Disinfection by Ozone, Ozone.
of
Sci. Eng. 16 (1994) 455−473.
ro
[17] P.B. Clancy, Treatment of Chlorinated Ethenes in Groundwater with Ozone and Hydrogen
static mixer for mixing chemicals in ballast water treatment systems, J. Water Process Eng. 8
na
(2015) 209−220.
[19] A. Cybulski, K. Werner, Static mixers-criteria for application and selection, Int. Chem. Eng.
ur
26 (1986) 171−180.
Jo
[20] J.P. Gingras, L. Fradette, P. Tanguy, J. Bousquet, Inline Bitumen Emulsification Using
[21] A.S. Klett, A.M. Payne, M.C. Thies, Continuous-Flow Process for the Purification and
Fractionation of Alkali and Organosolv Lignins, ACS Sustain. Chem. Eng. 4 (2016) 6689−6694.
[22] M. Rauline, P.A. Tanguy, J.M. Le Blevec, J. Bousquet, Numerical Investigation of the
37
Journal Pre-proof
[23] E.B. Nauman, B.A. Buffham, Mixing in Continuous Flow Systems, Chem. Eng. Commun.
29 (1983) 369−370.
[24] I. Manas-Zloczower, Studies of Mixing Efficiency in Batch and Continuous Mixers, Rubber
[25] S. Shojaee, S.H. Hosseini, A. Rafati, G. Ahmadi, Prediction of the Effective Area in
Structured Packings by Computational Fluid Dynamics, Ind. Eng. Chem. Res. 50 (2011)
of
10833−10842.
ro
[26] S.H. Hosseini, S. Shojaee, G. Ahmadi, M. Zivdar, Computational fluid dynamics studies of
-p
dry and wet pressure drops in structured packings, J. Ind. Eng. Chem. 18 (2012) 1465−1473.
re
[27] S. Liu, A.N. Hrymak, P.E. Wood, Laminar mixing of shear thinning fluids in a SMX static
lP
[28] A. Baumann, S.A.K. Jeelani, B. Holenstein, P. Stossel, E.J. Windhab, Flow regimes and
drop break-up in SMX and packed bed static mixers, Chem. Eng. Sci. 73 (2012) 354−365.
ur
[29] J.M. Zalc, S.E. Szalai, F.J. Muzzio, S. Jaffer, Characterization of Flow and Mixing in an
Jo
[30] J.M. Zalc, S.E. Szalai, F.J. Muzzio, Mixing Dynamics in the SMX Static Mixer as a
Function of Injection Location and Flow Ratio, Polym. Eng. Sci. 43 (2003) 875−890.
[31] D.M. Hobbs, P.D. Swanson, F.J. Muzzio, Numerical characterization of low Reynolds
number flow in the Kenics static mixer, Chem. Eng. Sci. 53 (1998) 1565−1584.
38
Journal Pre-proof
[32] O. Byrde, M.L. Sawley, Optimization of a Kenics static mixer for non-creeping flow
[33] E. Saatdjian, A.J.S. Rodrigo, J.P.B. Mota, On chaotic advection in a static mixer, Chem.
[34] M. Regner, K. Östergren, C. Träġardh, Effects of geometry and flow rate on secondary flow
and the mixing process in static mixers−A numerical study, Chem. Eng. Sci. 61 (2006)
of
6133−6141.
ro
[35] H.S. Song, S.P. Han, A general correlation for pressure drop in a Kenics static mixer, Chem.
[37] P. Joshi, K.D.P. Nigam, E.B. Nauman, The Kenics static mixer: new data and proposed
[38] M. Rafiee, S. Bakalis, P.J. Fryer, A. Ingram, Study of laminar mixing in kenics static mixer
Jo
678−684.
[39] M.K. Singh, P.D. Anderson, H.E.H. Meijer, Understanding and Optimizing the SMX Static
[40] S.S. Soman, C.M.R. Madhuranthakam, Effects of internal geometry modifications on the
dispersive and distributive mixing in static mixers, Chem. Eng. Process. 122 (2017) 31−43.
39
Journal Pre-proof
[41] H.-B. Meng, M.-Y. Song, Y.-F. Yu, X.-H. Jiang, Z.-Y. Wang, J.-H. Wu, Enhancement of
Laminar Flow and Mixing Performance in a Lightnin Static Mixer, Int. J. Chem. React. Eng. 15
(2016) 20160112.
[42] H. Meng, F. Wang, Y. Yu, M. Song, J. Wu, A Numerical Study of Mixing Performance of
High-Viscosity Fluid in Novel Static Mixers with Multitwisted Leaves, Ind. Eng. Chem. Res. 53
(2014) 4084−4095.
of
[43] H. Meng, J. Zhu, Y. Yu, Z. Wang, J. Wu, The effect of symmetrical perforated holes on the
ro
turbulent heat transfer in the static mixer with modified Kenics segments, Int. J. Heat Mass
reactive mixing in a novel coaxial jet static mixer, Chem. Eng. Process. 122 (2017) 190−203.
lP
mixer to achieve an energy-efficient mixing operation of non-newtonian fluids, 124 (2018) 1−10.
ur
static mixers: process intensification perspective. Rev. Chem. Eng. 2019, doi:10.1515/revce-
2017-0104.
Newtonian and non-Newtonian fluids in SMX mixers using PEPT, Chem. Eng. Res. Des. 108
(2016) 126-138.
40
Journal Pre-proof
Turbulent Drop Breakage in a Kenics Static Mixer and Comparison with Experimental Data,
[49] M.M. Haddadi, S.H. Hosseini, D. Rashtchian, G. Ahmadi, CFD Modeling of Immiscible
Liquids Turbulent Dispersion in Kenics Static Mixers: Focusing on Droplet Behavior, Chines J.
of
[50] M. Al-Atabi, Design and Assessment of a Novel Static Mixer, Can. J. Chem. Eng. 89 (2011)
ro
550−554.
-p
[51] J. Sir, Z. Lecjaks, Pressure Drop and Homogenization Efficiency of a Motionless Mixer,
re
Chem. Eng. Commun. 16 (1982) 325−334.
lP
[52] R.B. Bird, W.E. Stewart, E.N. Lightfoot, Transport Phenomena, New York. John Wiley &
Sons. (2002).
na
Corrugated Static Mixers for Turbulent Applications, Ind. Eng. Chem. Res. 51 (2012)
Jo
15986−15996.
[54] V. Kumar, V. Shirke, K.D.P. Nigam, Performance of Kenics static mixer over a wide range
[55] W.L. Wilkinson, M.J. Cliff, An Investigation into the Performance of a Static In-Line
[56] N.F. Shah, D.D. Kale, Two-Phase Gas-Liquid Flows in Static Mixers, AIChE J. 38 (1992)
308−310.
41
Journal Pre-proof
[57] C.D. Grace, Static Mixing and Heat Transfer, Chem. Process. Eng. 52 (1971) 57−59.
[58] S.H. Hosseini, G. Ahmadi, M. Olazar, CFD simulation of cylindrical spouted beds by the
[60] M.D. Das, A.N. Hrymak, M.H.I. Baird, Laminar liquid-liquid dispersion in the SMX static
of
mixer, Chem. Eng. Sci. 101 (2013) 329−344.
ro
Captions: -p
re
Table 1. Geometrical parameters of the static mixers.
lP
Table 3. Grid size, CFD simulation time, number of elements, estimated Z and CoV for the
Table 4. Grid size, CFD simulation time, number of elements, estimated Z and CoV for the
Jo
Table 5. Grid size, CFD simulation time, number of elements, estimated Z and CoV for the SMX
mixer at Re=160
Table 6. Grid size, CFD simulation time, number of elements, estimated Z and CoV for the new
mixer at Re=160
42