Download as pdf or txt
Download as pdf or txt
You are on page 1of 5

DESIGN AND ASSESSMENT OF A NOVEL STATIC MIXER

Mushtak Al-Atabi*
School of Engineering, Taylor’s University College, No 1, Jalan Taylor’s, 47500 Subang Jaya, Selangor DE, Malaysia

This paper examines the performance of a novel static mixer comprising a circular tube fitted with eight alternating equi-spaced semicircular rigid
insert (baffles) as the mixer elements. Experiments were carried out to obtain the coefficient of variance (CoV) for the mixing of two streams of
water and brine for Reynolds number between 60 and 700. Decreasing the baffles clearance ratio significantly reduces the CoV but at a cost of
an increase in the pressure drop across the static mixer. The presence of the mixing elements (baffles) promotes a non-laminar, turbulent-like
flow which considerably enhances the mixing. The static mixer described here represents a cost effective, easy to manufacture, low maintenance,
and flexible alternative to the more sophisticated static mixers currently in use.

Ce document examine le rendement d’un nouveau mélangeur statique composé d’un tube circulaire comportant huit garnitures rigides demi-
circulaires équidistantes alternantes (déflecteurs) comme éléments du mélangeur. Des expériences ont été réalisées pour obtenir le cœfficient
de variation (CV) pour le mélange de deux jets d’eau et de saumure pour un nombre de Reynolds entre 60 et 700. Diminuer le rapport de
dégagement des déflecteurs réduit considérablement le CV, mais il en coûte une augmentation de la perte de charge dans tout le mélangeur
statique. La présence des éléments de mélange (déflecteurs) favorise un écoulement non laminaire semblable à de la turbulence qui accroı̂t
considérablement le mélange. Le mélangeur statique décrit ici représente une solution de rechange souple, économique, facile à fabriquer et
présentant peu d’entretien aux mélangeurs statiques plus élaborés utilisés à l’heure actuelle.

Keywords: static mixer, homogenisation, segmental baffles, coefficient of variance, laminar flow, fluid mechanics

INTRODUCTION and reduced maintenance requirements compared with mechan-


ically stirred mixers. They offer a more controlled and scaleable

M
ixing is an important process in many industries. It is rate of dilution in fed batch systems and can provide homogeni-
usually achieved by using either mechanically stirred sation of feed streams with a minimum residence time. They also
vessels or static mixers. Static mixers consist of a series provide good mixing at low shear rates where locally high shear
of stationary and rigid inserts as the mixing element installed in rates in a mechanical agitator may damage sensitive materials.
pipes, ducts, or transfer tubes. The purpose of these inserts is to Static mixers also come with self-cleaning features. Interchange-
redistribute fluid in a direction transverse to the direction of the able and disposable static mixers are also available (Etchells and
main flow, both tangentially and radially. Static mixers find appli- Meyer, 2004).
cations in wide variety of processes such as blending of miscible Numerous static mixer designs have been proposed. Thakur
fluids both in laminar and turbulent flows, mixing and disper- et al. (2003) reported the existence of approximately 2000 US
sion of immiscible fluids by helping to generate an interface, solid patents and more than 8000 literature articles that describe static
blending, heat and mass transfer and homogenisation (Etchells mixers and their applications. More than 30 models have made
and Meyer, 2004). These processes, in turn, serve many industries their way into different industries and they are commercially
such as chemical and agricultural chemicals production, grain available. Kenics mixers, KMS and KMX, (by Chemineer, Inc.,
processing, food processing, minerals processing, petrochemical Dayton, OH) are used in both laminar and turbulent flow regimes
and refining, pharmaceuticals and cosmetics, polymers, plastics,
textiles, paints, resins and adhesives, pulp and paper, water and
waste water treatment. ∗ Author to whom correspondence may be addressed.
The use of static mixers in continuous processes is an attractive E-mail address: mushtak.t@taylors.edu.my
alternative to conventional agitation since similar and sometimes Can. J. Chem. Eng. 89:550–554, 2011
better performance can be achieved at lower cost. As they have © 2010 Canadian Society for Chemical Engineering
DOI 10.1002/cjce.20412
no moving parts, stationary mixers typically have lower energy Published online 9 November 2010 in Wiley Online Library
consumptions, smaller space requirements, low equipment cost (wileyonlinelibrary.com).

| 550 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 89, JUNE 2011 |
Figure 1. Geometries of some commercially available mixers.

to achieve thermal homogenisation of polymer melt (Chen, 1975), mixers (Bertrand et al., 1994; Lang et al., 1995; Hobbs et al., 1998;
gas–liquid dispersion and dilution of feed to reactor (Smith, 1978) Rauline et al., 1998; Visser et al., 1999).
and dispersion of viscous liquids (Berkman and Calabrese, 1988). This paper reports an experimental investigation of the mixing
Another mixer manufactured by Chemineer, Inc., is the HEV (high of two streams of water and brine (25% NaCl, w/w) in a static
efficiency vortex) static mixer which is used for turbulent low vis- mixer that employs staggered baffles as the mixing elements in
cosity liquid–liquid blending and gas–gas mixing (Thakur et al., a circular tube. The quality of the mixing is inferred from the
2003). degree of mixedness between the two streams as exhibited by the
Koch-Glitsch (Wichita, KS), produces a range of static mixers distribution of the concentration of NaCl at a cross sectional plane
that comprise the SMX, SMV, SMXL, SMF, and SMR (Thakur et downstream of the static mixer.
al., 2003). SMX mixers are used for laminar mixing of high vis-
cosity liquids and liquids with extremely diverse viscosity and
homogenisation of melts in polymer processing. SMV mixers find MATERIALS AND METHODS
applications in turbulent low viscosity mixing and mass trans-
fer in gas–liquid systems. SMXL static mixers are used in the Static Mixer Design
laminar flow regime for liquid–liquid extraction, homogeneous The static mixer presented in this paper comprises a circular tube
dispersion, and laminar heat transfer enhancement for viscous with semicircular staggered rigid inserts (baffles). Similar geome-
fluids. SMF mixers are utilised for sludge conditioning, pulp stock tries are often encountered in shell and tube heat exchangers
blending, bleaching, dilution, bleaching of suspension and slur- (Kakac and Liu, 2002) and have recently been employed as an
ries. SMR mixers are used for laminar polystyrene polymerisation idealised biological flow device (Al-Atabi et al., 2004). The use
and devolatilisation. Other mixers produced by different manufac- of this configuration as a static mixers potentially offers a set of
turers are also available. Some of the available static mixers are benefits; it is easy to manufacture and install, can be adapted
shown in Figure 1. to non-circular ducts, offers more flexibility in changing design
The assessment of the performance of a static mixer is based parameters (baffle spacing and clearance ratio) and it can offer a
usually on two criteria: the quality of mixing and the pressure bi-directional operation.
drop through the mixer. Mixing quality is often ascertained by Figure 2 shows a schematic diagram of the static mixers
visual observations which may be carried out by simple dye designed and analysed in this study. The mixers are made of Plex-
addition or blending and curing two colour epoxies. Quantita- iglas pipe with an internal diameter, D, of 21 mm fitted with eight
tive data may be obtained from determining the coefficient of equi-spaced baffles, b/D = 1.5, where b is the baffle spacing. Three
variation based on various statistical analysis of mixing quality models were constructed with clearance ratios, c/D, of 0.3, 0.5,
(Myers et al., 1997). Despite the numerous applications of static and 0.7, where c is the clearance at the baffle section. The flow
mixers, the fundamental understanding of the process is so far in a tube with no baffle (c/D = 1.0) was studied as well for rig
insufficiently understood (Godfrey, 1992). Some researchers have validation and for comparison.
reported pressure drop data (Shah and Kale, 1992; Li et al., 1997; Flow visualisation studies (Al-Atabi et al., 2005) have suggested
Thakur et al., 2003), residence time distributions (Nauman, 1991; that the segmental baffles fitted in tubes encourage mixing. Figure
Shah and Kale, 1992), friction factor (Legrand et al., 2001), or 3a adapted from Al-Atabi et al. (2005) shows typical qualitative
mixing pattern “frieze” by solidification of coloured components results of these visualisation studies. The mixing achieved using
(Etchells and Meyer, 2004). Over the years computational sim- this rather simple configuration is similar to the mixing achieved
ulations of mixing analysis have been carried out for different using more established (and sophisticated) static mixers such as

Figure 2. Cross-sectional (A–A) view of the development length and the static mixer.

| VOLUME 89, JUNE 2011 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 551 |
Figure 3. Flow structure at Reynolds number = 150. Flow from left to right. (a) The staggered baffles static mixer c/D = 0.3 (Al-Atabi et al., 2005). (b)
SMX static mixer (four mixing elements) (Al-Atabi and Abakr, 2007).

Figure 4. The multi-point sampling probe. (a) Front view. (b) Cross-sectional slide view.

SMX as shown in Figure 3b (Al-Atabi et al., 2005), while main-


taining a lower pressure drop penalty. The present paper aims to
provide a quantitative understanding of the degree of mixing and
the accompanying pressure drop in tubes with segmental baffles.

Test Facility and Procedure


The mixers performance is assessed by the degree of mixedness
between two streams of water and brine (25% NaCl, w/w). Both
the brine and water are supplied from their respective reservoirs
into constant head elevated tanks using two separate pumps. The
liquid levels in both tanks are kept the same. Water is continu-
ously supplied to the static mixer and its flow rate is controlled by
a downstream valve. Brine is injected via a 3 mm diameter tube in
the geometric centre of the tube 3D upstream of the static mixer.
A multi-point sampling probe has been constructed to extract
nine mixture samples simultaneously from nine points evenly
distributed across the pipe cross-section 3D downstream of the
mixer. The cross sectional area of each probe is less than 0.5%
of the cross-sectional area of the mixer, hence it will have mini-
mum disturbance to the flow. Figure 4a shows the front view of Figure 5. Experimental set-up.
the probe and Figure 4b its cross-sectional view. A schematic of
the test rig is given in Figure 5.
Refraction index (RI) of the sample was measured using a
refractometer in order to infer the average NaCl concentration in RESULTS
the sample. Pressure tappings were located 1.5D upstream of the
first baffle and 1.5D downstream of the last baffle. An inclined Mixing Quality
water manometer recorded pressure difference across the static The refractometer was calibrated with known concentrations, of
mixer. The Reynolds number, based on the tube diameter and NaCl and a linear relationship was established.
total flow rate, was varied from 50 to 6000 for each of the three The degree of mixing is defined by the coefficient of vari-
mixers and the tube with no baffle. ance (CoV), Equation (1), downstream of the mixer (Etchells and

| 552 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 89, JUNE 2011 |
Figure 6. Variation of coefficient of variance (CoV) with Reynolds
number (Re). Figure 7. Effect of clearance ratio on the friction factor.

Meyer, 2004):

CoV = (1)

where  is the standard deviation and x̄ the mean concentration
of brine samples taken from all ports in an experiment run under
a given set of conditions. A CoV value of zero indicates perfect
mixing, conversely a value of unity means no mixing. In a typical
industrial mixing process, an additive is considered well mixed at
5% CoV (Etchells and Meyer, 2004).
Figure 6 shows the effect of baffle configuration on CoV for
50 < Re < 700. For the tube with no baffle (c/D = 1.0) CoV is inde-
pendent of Re as the flow at these conditions was clearly laminar;
there is little mixing between streamlines in such flow and molec-
ular diffusion is an inefficient mixing process. For tubes with Figure 8. Variation of (CoV/f) ratio with the Reynolds number.
baffles, increasing the Reynolds number decreases the CoV; the
larger the c/D, the greater the decrease in CoV. However, at a
given Re, the CoV is less for a smaller c/D. Hence, at Re = 50, the increases. Similar behaviour is noted for the c/D = 0.7. However,
tube with c/D = 0.3 shows the least CoV and the greatest degree this trend reverses for c/D = 0.5 and 0.3.
of mixing. The experimental error varied from ±5 to ±10% for
CoV at a given Reynolds number.
DISCUSSION
Friction Factors The laminar flow condition is well known to provide the worst
Figure 7 shows the effects of the baffles on the friction factor case for mixing. This is clearly seen in Figure 6 where the CoV
for 50 < Re < 5000 and compares those with some available static for the tube without baffles remains constant for 50 < Re < 700.
mixers data and circular tubes data. The friction factors of the The use of a series of equi-spaced semicircular inserts in the tube
proposed static mixers are higher than those of circular tubes can significantly improve mixing even in nominally laminar flow
but generally are lower than those of KMX (Reeder et al., 2001) conditions. This is the basis of the proposed novel static mixer
and Kenics mixers (Joshi et al., 1995). The general definition of design. Figure 6 also shows, that for the same number of baffles,
friction factor is given by Equation (2). Theoretical friction fac- at a given Reynolds number, decreasing the clearance of the baf-
tors in circular tubes are given by f = 64/Re for laminar flow and fles decreases the CoV and hence improves the mixing. Hence at
f = 0.079/Re0.25 for turbulent flow: Re = 100, decreasing the clearance from 70% of the tube diameter
to 30% improves the mixing by a factor of 4. At Re = 700, the CoV
P D decreases by a factor of 10 and mixing improves by an order of
f = (2)
1
2
L
V 2 magnitude over the same decrease in the clearance ratio. For a
given baffle configuration, CoV decreases with increasing Re. For
where P is the pressure drop over the length of the test section, c/D = 0.7, CoV drops from 0.4 to 0.2 when Re increases from 100
L, D the tube diameter, V the mean velocity defined by the volume to 700; but for c/D = 0.3, the CoV decreases by a factor of 10.
flow rate Q, and  the density of water as an approximation of the Flow visualisation studies, such as those shown in Figure 3,
density of the brine water mixture. show that non-laminar, turbulent-like flow conditions are cre-
Figure 8 attempts to combine the effect of Reynolds number on ated by the baffles even when the flow upstream of the baffles
both the CoV and friction factor for design optimisation purposes. is clearly laminar. It would require hot-wire or LDA to detect any
For any Re, the value of CoV/f increases with the increase of the fluctuations in flow to determine if the flow had undergone tran-
baffle clearance. For c/D = 1.0, the CoV tends to increase as the Re sition to turbulence. It is well known that transition can occur at

| VOLUME 89, JUNE 2011 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 553 |
a Reynolds number as low as 10 in porous media (Bear, 1988). Paul, V. A. Atiemo-Obeng and S. M. Kresta, Eds.,
The introduction of obstacles to flow is expected to increase the Wiley-Interscience, New Jersey (2004).
pressure difference required to drive the flow at a given flow rate. Godfrey, J. C., “Static Mixers,” in “Mixing in the Process
Figure 7 compares the friction factor, that is, pressure drop, in Industries,” 2nd edition, Chap. 12, N. Harnby, M. F. Edward
circular tubes and commercial static mixers with those in tubes and A. W. Nienow, Butterworth-Heinemann, Oxford (1992).
with various baffle configurations. At a given Re, the friction fac- Hobbs, D. M., P. D. Swanson and F. J. Muzzio, “Numerical
tor increases with decreasing clearance. However, it is interesting Characterization of Low Reynolds Number Flow in the Kenics
to note that unlike Poiseuille flow, the friction factor for each baf- Static Mixer,” Chem. Eng. Sci. 53, 1565–1584 (1998).
fle configuration at first decreases with increasing Re but soon Joshi, P., K. D. P. Nigam and E. B. Nauman, “The Kenics Static
becomes independent of Reynolds number. For c/D = 0.7, that Mixer: New Data and Proposed Correlations,” Chem. Eng. J.
occurs at Re ≈ 1000, at c/D = 0.3, it occurs at Re ≈ 400. It is a 59, 265–271 (1995).
characteristic of fully developed turbulent flow that the friction Kakac, S. and H. Liu, “Heat Exchangers Selection, Rating, and
factor depends only on surface roughness and not on Reynolds Thermal Design,” 2nd edition, CRC Press LLC, Florida (2002).
number. Perhaps the baffles induce much earlier onset of tran- Lang, E., P. Drtina, F. Streiff and M. Fleischli, “Numerical
sition/turbulence in tube with baffles. When Cov/f is plotted Simulation of the Fluid Flow and Mixing Process in a Static
against Re, it can be seen that the best mixing with the mini- Mixer,” Int. J. Heat Mass Transfer 38, 2239–2250 (1995).
mum pressure loss occurs with a c/D of 0.3. This is particularly Legrand, J., P. Morançais and G. Carnelle, “Liquid–Liquid
evident for Reynolds numbers about 450. As well as the superior Dispersion in an SMX-Sulzer Static Mixer,” 4th International
mixing for a given pressure drop, it is clear that this design of Symposium on Mixing in Industrial Processes 79, 949–956
baffle mixer will also produce a more compact device. (2001).
Li, H. Z., C. Fasol and L. Choplin, “Pressure Drop of Newtonian
and Non-Newtonian Fluids Across a Sulzer SMX Static
CONCLUSIONS Mixer,” Trans. IChemE 75A, 792–796 (1997).
Semicircular rigid inserts were introduced into circular tubes as Myers, K., A. Bakker and D. Ryan, “Avoid Agitation by Selecting
the mixer elements in a novel static mixer. The effectiveness of Static Mixers,” Chem. Eng. Prog. 28–37 (1997).
this design has been found from measurement of CoV of brine Nauman, E. B., “On Residence Time and Trajectory Calculations
in the mixing of two laminar streams of brine and water in three in Motionless Mixers,” Chem. Eng. J. 47, 141–148 (1991).
configurations of baffles in a straight tube. In all cases, CoV values Rauline, D., P. A. Tanguy, J. M. Le Blevec and J. Bousquet,
show that the baffles have promoted mixing. Flow visualisation “Numerical Investigation of the Performance of Several Static
studies suggest that the baffles induce non-laminar turbulent like Mixers,” 76(3), 527–535 (1998).
flow which aids mixing. The increase in mixing is accompanied Reeder, M. F., J. Fasano and K. Myers, “Characterization of
by an increase in pressure drop as shown by the increase in fric- blending in static mixers,” Proc. of Mixing XVIII,
tion factor in all cases when compared with similar Poiseuille Pennsylvania, USA, June 24–29 (2001).
flow. A parametric study needs to be carried out to optimise the Shah, N. F. and D. D. Kale, “Two-Phase, Gas–Liquid Flows in
configurations of the baffles, for example, number of baffles and Static Mixers,” AIChE J. 38, 308–310 (1992).
baffle spacing, for maximum mixing. It is envisaged that such a Smith, J. M., “Gas Dispersion in Viscous Liquids With a Static
mixer may find applications in mixing solids, very viscous liquids Mixer,” Chem. Eng. 827–830 (1978).
or biofluids which might be damaged by the high strain rates in Thakur, R. K., C. Vial, K. D. P. Nigam, E. B. Nauman and G.
turbulent or mechanical mixing conditions. Djelveh, “Static Mixers in the Process Industries—A Review,”
Trans. IChemE 81, 787–826 (2003).
Visser, E. J., P. F. Rozendal, H. W. Hoogstraten and A. C. M.
REFERENCES Beenackers, “Three Dimensional Numerical Simulation of
Al-Atabi, M. T., S. B. Chin and X. Y. Luo, “An experimental Flow and Heat Transfer in the Sulzer SMX Static Mixer,”
study of the flow in an idealized human cystic duct,” in Proc. Chem. Eng. Sci. 54, 2491–2500 (1999).
of the First Asian Pacific Biomechanics Conference, Osaka,
Japan, March 25–28 (2004).
Al-Atabi, M. T., S. B. Chin and X. Y. Luo, “Flow Structure in Manuscript received February 19, 2010; revised manuscript
Circular Tubes With Segmental Baffles,” J. Flow Vis. Image received April 21, 2010; accepted for publication April 30, 2010.
Process. 8(2), 89 (2005).
Al-Atabi, M. T. and Y. A. Abakr, “Laminar Mixing in SMX Static
Mixers,” J. Eng. Sci. Technol. 2(1), 95–101 (2007).
Bear, J., “Dynamics of Fluid Flow in Porous Media,” Dover
Publications, Inc., New York (1988).
Berkman, P. D. and R. V. Calabrese, “Dispersion of Viscous
Liquids by Turbulent Flow in a Static Mixer,” AIChE J. 34,
602–609 (1988).
Bertrand, F., P. A. Tanguy and F. Thibault, “A Numerical Study
of the Residence Time Distribution in Static Mixing,” IChemE
Symp. Ser. 136, 163 (1994).
Chen, S. J., “Static Mixing of Polymers,” Chem. Eng. Prog. 71,
80–883 (1975).
Etchells, A. W. III and C. F. Meyer, “Mixing in Pipelines,” in:
“Handbook of Industrial Mixing Science and Practice,” E. L.

| 554 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 89, JUNE 2011 |

You might also like