You are on page 1of 37

Mixing time in stirred vessels: A review of experimental techniques

Gabriel Ascanio

PII: S1004-9541(15)00087-7
DOI: doi: 10.1016/j.cjche.2014.10.022
Reference: CJCHE 246

To appear in:

Received date: 13 December 2013


Revised date: 2 July 2014
Accepted date: 28 October 2014

Please cite this article as: Gabriel Ascanio, Mixing time in stirred vessels: A review of experimental
techniques, (2015), doi: 10.1016/j.cjche.2014.10.022

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

2013-0569
搅拌容器中的混合时间:实验技术的综述

Graphic abstract

Mixing time is defined as the time required for achieving a certain degree of
homogeneity of injected tracer in a unit operation vessel. It has been used as a key
parameter for assessing the performance of a mixing system. From an experimental

MANUSC
standpoint, several techniques have been developed for measuring the mixing time.
Based on the disturbances to flow, they can be classified into two groups: non-intrusive
and intrusive. A review of the experimental techniques reported in the literature in last

RIPT
50 years for the measurement of mixing time in stirred vessels under single and gas-
liquid flow conditions with Newtonian and non-Newtonian fluids in the laminar and
turbulent regime is made, and a comparison between these techniques is also presented.

1
ACCEPTED MANUSCRIPT

Fluid Dynamics and Transport Phenomena


Mixing time in stirred vessels: A review of experimental
techniques*

Gabriel Ascanio**
Center of Applied Sciences and Technological Development, National Autonomous University of
Mexico, Circuito Exterior, Ciudad Universitaria, DF 04510 Mexico***

Article history:
Received 13 December 2013
Received in revised form 2 July 2014
Accepted 28 October 2014

*Supported by DGAPA-UNAM through the grant IN-108312.


** Corresponding author. Email: gabriel.ascanio@ccadet.unam.mx
***Centro de Ciencias Aplicadas y Desarrollo Tecnológico, Universidad Nacional Autónoma de
México, Circuito Exterior, Ciudad Universitaria, DF 04510 Mexico

Abstract Mixing time is defined as the time required for achieving a certain degree of
homogeneity of injected tracer in a unit operation vessel. It has been used as a key
parameter for assessing the performance of a mixing system. From an experimental
standpoint, several techniques have been developed for measuring the mixing time.
Based on the disturbances to flow, they can be classified into two groups: non-intrusive
and intrusive. However, depending on the type of data generated, they can be also
classified into direct measurements and indirect measurements (Eulerian and
Lagrangian). Since the techniques available for measuring mixing times in an agitated
tank do not provide the same information, its choice depends on several factors, namely:
accuracy, reproducibility, suitability, cost, sampling speed, type of data, and processing
time. A review of the experimental techniques reported in the literature in last 50 years
for the measurement of mixing time in stirred vessels under single and gas-liquid flow
conditions with Newtonian and non-Newtonian fluids in the laminar and turbulent
regime is made, and a comparison between these techniques is also presented.
Keywords mixing time, stirred vessel, homogeneity.

2
ACCEPTED MANUSCRIPT

1 Introduction
Mixing time is a key parameter used for analyzing the performance and the
hydrodynamics of a stirred vessel. In general terms, the mixing time is defined as the
time required for achieving a certain degree of homogeneity of tracer inserted in a
stirred vessel [1]. From a macromixing standpoint, the bulk mixing time is the time
required to get all points in the vessel uniformly distributed, while the local mixing time
is the measure of how fast a material is distributed in a particular region of the vessel,
which depends on the local turbulence.

Therefore, local measurements are time and space dependent, while bulk mixing time
are based on time dependent measurements (temporal measurements) [2]. Mixing time
can be expressed in its nondimensional form:

N m  K (1)

where m is the mixing time in s, N the impeller speed in revolutions per second and K
is a constant which depends on the size, geometry of the tank and the flow regime [3].
Constancy of K is valid in the laminar and turbulent regimes but not in the transitional
one. Moo-Young et al. [4], the mixing time can be correlated in a dimensionless manner
as
N  m   Re  (2)

where  and  are adjustable parameters and Re is the Reynolds number,


ND2
Re   (3)

where  is the fluid density in kg/m3, D is the impeller diameter in m and µ is the fluid
dynamic viscosity in Pa·s.
In the case of non-Newtonian fluids, such a number is defined by
 N 2n D 2
Re  (4)
k
where k and n are the consistency index and the flow behavior index, respectively, for
fluids obeying the Ostwald-de Waele model (power-law model), in which the rate of
deformation has been taken as   N . However, the Reynolds number for non-
Newtonian can be also defined in terms of the effective viscosity, which is a function of
the shear rate. Nowadays, the most used definition of the shear rate in stirred vessels is

3
ACCEPTED MANUSCRIPT

the one developed by Metzner and Otto [5], which is calculated from the power-law
model:
1/n1
  ks N
   k  (5)
 
where the shear rate is proportional to the impeller rotational speed, being ks the
proportionality constant. Based on this approach, the Reynolds number is defined as
 N D2  N 2n  D2
Re   (6)
 kk s n1
A number of mixing time correlations have been reported in the literature taking
into account the effect of baffles on mixing time [6-8], as well as the relationship
between circulation and mixing times [9]. In the turbulent regime, the dimensionless
mixing time is independent on the Reynolds number and it can be correlated by
empirical expressions involving the size and geometry of the tank as well as some
hydrodynamic parameters [10,11]. Experimental techniques for measuring mixing time
can be classified depending basically on two different scenarios, namely: 1) level of
disturbance to flow, 2) type of data collected. Based on the first scenario, the techniques
are classified as non-intrusive and intrusive. With respect to the intrusive techniques,
they are based on local and global measurements. Although they are accurate, its use is
limited because the flow pattern is modified by the presence of probes in the vessel.
Mavros [12] reviewed the experimental techniques available for the study of flow
patterns in stirred vessels. Although the aim of his paper was to give a general overview
of intrusive and non-intrusive techniques for the flow visualization in tanks, some of
them can be also applied also for estimating mixing times.
On the other hand, considering the type of data collected, these techniques are
classified as direct measurements or indirect measurements (Eulerian data or
Lagrangian data). In the case of indirect measurements, mixing time is to be inferred
from physical measurements of conductivity or velocity for instance, while direct
techniques provide real time results or at least requiring minimum processing. Table 1
summarizes the classification of experimental techniques as a function of both
scenarios.

Table 1 Classification of measurement techniques of mixing times in stirred vessels

Flow disturbance

4
ACCEPTED MANUSCRIPT

Non-intrusive Intrusive
 Colorimetry  Conductometry and pH (probes)
 Electrical resistance tomography
 Positron emission particle tracking
 Planar laser induced fluorescence
 Thermography

Type of data provided

Indirect measurement
Direct measurement
Eulerian Lagrangian
 Colorimetry  Thermography  Electrical resistance
 Planar laser induced tomography
fluorescence
 Conductometry and
pH (probes)

Nere et al. [8] performed a critical review of the literature dealing with liquid-phase
mixing in stirred vessels in the turbulent regime. Although, the limitations of the
techniques employed for measuring mixing times were described therein, they focused
on the effects of different parameters such as impeller geometry, impeller position on
liquid-phase mixing as well as the mathematical models developed for analyzing the
hydrodynamics in stirred vessels. In this paper, the experimental techniques reported in
the literature over last 50 years for the measurement of mixing time in stirred vessels
have been reviewed by highlighting their advantages and drawbacks along with the
most outstanding findings. This review covers single phase and gas-liquid mixing of
Newtonian and non-Newtonian fluids in the laminar and turbulent regimes.

2 EXPERIMENTAL TECHNIQUES
2.1 Colorimetry
From a practical standpoint, colorimetry is by far the most common technique
employed for measuring mixing times in stirred vessels [13-15]. It is a non-intrusive
technique extensively reported in the literature, which is used not only to determine the
time required to achieve the desired degree of homogenization, but also to visualize
qualitatively flow patterns and to reveal the presence of secondary flows generated

5
ACCEPTED MANUSCRIPT

under steady stirring such as well-mixed regions (caverns), islands and other segregated
regions like stagnant of dead flow zones [16-20]. The technique basically consists of
injecting a liquid tracer and observing how it is dispersed in the fluid contained in
stirred vessels. A variant of this technique is based on the color-decolorization approach
by using a pH sensitive tracer. The reader is referred to Ascanio et al. [21] for a detailed
description of the measuring protocol. Figure 1 shows an example of a typical mixing
sequence by the colorimetric technique.

Figure 1 Typical mixing evolution in a stirred tank using colorimetry [21].

MANUSC
Norwood and Metzner [22] reported for the first time the use of an acid-base

RIPT
neutralization reaction for measuring mixing time in baffled stirred tanks equipped with
turbines operating in the turbulent regime, which has been extensively adopted in the
research. Hari-Prajitno et al. [23] used also this technique for measuring mixing time
under aerated and unaerated conditions for dual and triple coaxial impeller
configurations in the turbulent regime. The decolorization method has been not only
used for measuring mixing time, but also for getting the insight of flow patterns
generated by different impeller such as close-clearance impellers, multiple and coaxial
arrangements, inclined impeller, etc. [24-34].
Although colorimetry is a very simple technique to implement, mixing time
estimation depends strongly on the subjectivity of the human observer. For that reason,
complementary techniques based on image processing have been developed [35-40].
Kouda et al. [35] dissolved phenolphthalein as pH indicator in bacterial cellulose and
CMC solution, which became dark purple when NaOH being present. Then, H 2SO4 was
added at time zero and the images were captured every 0.5 s. Afterwards, the digital
brightness of images was analyzed by image processing, and the mixing time was
determined in terms of the decolorization ratio (output/reference).
Table 2 summarizes the operating principle, advantages and disadvantages of the
estimation of mixing time by colorimetry.

6
ACCEPTED MANUSCRIPT

Table 2 Operating principle, advantages and disadvantages of colorimetry


Operating principle Advantages Disadvantages
Color change of a  Direct measurement  Mixing time measurement
solution containing  Simple to implement depends on the human eye
a pH sensitive  Non-intrusive technique subjectivity without image
tracer  No calibration is processing
required  Temporal resolution with image
 Reveal flow structures processing techniques

 Used under aerated and  Spatial resolution with high-


unaerated conditions resolution pictures

 Accurate when  For transparent systems only.


combined with image  Limited applicability under
processing techniques highly aerated conditions in the
turbulent regime

2.2 Electrical resistance tomography


Electrical Resistance Tomography (ERT) is a non-intrusive and non-invasive
technique from which cross-sectional images showing the distribution of electrical
conductivity of gas-liquid flows in a stirred vessel from Lagrangian measurements taken
at the boundary of the vessel. It can be used where the continuous phase is conductive
and the other phases have different values of conductivity. A typical ERT setup consists
basically of a set of electrodes, the data acquisition system and a user interface [41]. The
array of sensors are usually a series of rings in which the electrodes are equally spaced
around the vessel circumference. Electrical measurements are made without affecting
the flow in the vessel. A typical measurement protocol consists of applying the current
to the next electrode pair and the resultant potential differences on all other electrode
pairs are measured. This step is repeated until all electrode pair combinations have been
used. A reconstruction technique converts the raw peripheral potential differences
measurements into a two-dimensional electrical conductivity distribution of the
measurement area (tomogram). Such a technique is based mainly on three different
algorithms: linear back projection (LBP), a modified Newton-Raphson method (MNR)
and a parametric model based technique (PM). The three algorithms use a single point

7
ACCEPTED MANUSCRIPT

calibration on a set of reference voltages from a sensor filled with a solution having a
known conductivity [42].
Figure 2 shows an example of a conductivity map showing the evolution of the
mixing process in the stirred vessel. For that purpose a color scale is used for showing
the variations in conductivity. For this example, blue regions indicate no mixing, green
regions partial mixing and red zone full mixing.

MANUSC
Figure 2 Conductivity maps in a stirred vessel [41].

RIPT
Although, ERT is a relatively new technology developed to get insight about the
hydrodynamics inside stirred vessels, among other applications, its usefulness in such
applications has been already reported [43-57] and validated with Computational Fluid
Dynamics (CFD) [58]. Wang et al. [48, 59] reported the use of an ERT system for the
gas-liquid mixing in a stirred vessel for discriminating the differences of three-
dimensional gas hold up variations in the liquid phase viscosity. Simultaneously, the
fluid dynamics was also investigated by following the spatiotemporal conductivity
changes of a fluid tracer injected in the working liquid. Figure 3 shows the real time
mixing curves of the liquid for the air-water mixing, which have been plotted from stack
of tomograms obtained in the bottom plane and in the top plane of the stirred vessel.
Looking at both curves, it is observed that the brine tracer is thoroughly mixed after 19 s
and from this point the concentration in the top and the bottom remains constant, which
can be confirmed by observing the corresponding tomograms.

8
ACCEPTED MANUSCRIPT

Figure 3 Liquid phase mixing curves for air-water mixing and the tomograms of the
liquid phase mixing of a brine tracer pulse (adapted from [48,59]).

MANUSC
Although most of the tomographic measurements have been done using sensors

RIPT
placed outside the tank, Abdullah et al. [60] reported the use of a series of conductivity
electrodes place into the tank to study the nature of gas-liquid-solid mixing in a gas-
induced stirred tank reactor. Edwards et al. [62] investigated the semibatch precipitation
of aluminum hydroxide by the addition of concentrated sodium hydroxide (NaOH) to
aluminum nitrate in a stirred vessel by using multimodal tomographic techniques. For
that purpose ERT was used to analyze the mixing performance during precipitation,
while positron emission particle tracking (PEPT) was used to study the flow field in the
vessel as a function of the fill level, power input, etc. They found by PEPT that mixing
time measured with ERT depends on the ratio of the feeding rate to the bulk velocity in
the feed region. PEPT is a useful technique for studying three-dimensional flow
phenomena in opaque systems [63]. The operating principle is based on a particle tracer
labeled with a radionuclide that decays via beta-plus decay generating two gamma rays.
A PEPT camera detects the two gamma rays and defines a line along which the
annihilation occurred. Fangary et al. [64] used the PEPT technique for determining the
flow patterns in water and in viscous non-Newtonian CMC solutions inside a vessel
agitated by axial flow impellers. Barigou et al. [65] validated the results obtained with a
PEPT system with PIV and LDA in a stirred vessel containing a transparent fluid.
Subsequently, the approach was applied to opaque systems containing low viscosity and

9
ACCEPTED MANUSCRIPT

high solid content suspensions obtaining the spatial distribution of suspended particles
of two sizes throughout the vessel as well as flow fields for liquid and solid phases.
Edwards et al. [62] combined ERT with PEPT to study the hydrodynamics of a stirred
vessel of aluminum hydroxide. In that case the temporal and spatial distribution of fed
ionic species provided by ERT was complemented by flow field information on the feed
plume provided by PEPT.
A variant of radioactive tracking consists of injecting a small volume of a
radioactive liquid tracer and monitoring its concentration. This technique offers some
advantages over optical methods such as the possibility of measuring over a wide
temperature range (more than 300 C) and working with nontransparent vessels [8].
The operating principle as well as the advantages and disadvantages of ERT
technique are summarized in Table 3.

Table 3 Operating principle, advantages and disadvantages of Electrical Resistance


Tomography (ERT)
Operating principle Advantage Disadvantage
Electrical measurements  Non-intrusive, non-  Expensive technique
of conductivity producing invasive  Calibration must be
cross-sectional images.  Series of 2-D tomograms made prior to
An image algorithm at different heights can be experiments
converts the raw used for producing 3-D  Time consuming
peripheral potential flow fields. assembly of setup
differences measurements  Suitable under aerated and  Good skills for
into a tomogram (two- unaerated conditions interpreting
dimensional electrical  Allows also observing measurements
conductivity distribution isolated mixing regions
of the measurement area).  Temporal and spatial
resolution
 For transparent and non-
transparent systems

10
ACCEPTED MANUSCRIPT

2.3 Planar laser-induced fluorescence


Planar laser-induced fluorescence (pLIF) is a non-intrusive Eulerian technique
suitable for the instantaneous measurement of concentration maps in liquid flows.
Figure 4 shows a pLIF setup, which consists basically of the following parts:
 Light source: Typically a Nd:YAG laser (neodymium-doped yttrium aluminum
garnet) or an argon-ion laser. Although a Nd:YAG laser emits light with a
wavelength of 1064 nm, it is also operated in both continuous or pulsed mode, in
which a high-intensity pulse can be generated at 532 nm or higher harmonics at
355 and 266 nm.
 During the experiments the dye is excited by laser light whose frequency closely
matches the excitation frequency of the dye. It absorbs the laser light energy at
short wavelength and re-emits light at a longer wavelength that can be detected
by a photodetector. The most common dyes used in pLIF are fluorescein and
rhodamine 6G.
 A charge-coupled device camera (CCD) with a narrow-band filter for capturing
fluorescent light only.

ACCEP
TED Figure 4 Basic setup of pLIF (adapted from [61])

Since the dye fluorescence is strongly dependent on the concentration, the pLIF
system must be firstly calibrated. If a dye is chosen for concentration measurements
(fluorescein or rhodamine 6G) the concentration signal (S) is calibrated with the
following [61]:
S  f optic Ac  EVc Q  C (7)

11
ACCEPTED MANUSCRIPT

where C is the dye concentration, E is the laser light intensity, Q is the quantum
efficiency of the dye (at the laser excitation wavelength), foptic corresponds to optical
factors, Vc is the sampling volume, AC represents the absorption phenomena integrated
on the light path (L) in the fluid characterized by the absorption index ( ), which is
calculated by
(8)
A  e  LC
c

For low concentration experiments, absorption phenomena are negligible, so that


AC = 1, leading to linear relationship between the signal (S) and (C,E,), being  a
constant that characterizes all experimental parameters. In such conditions, the
concentration is accurately measured as the amount of light received by the detector.
Many commercial LIF systems include a software capable of performing a quick
calibration of concentration, as well as signal corrections and other operations.
Once the system has been calibrated, the signals are processed by converting the
fluorescence images into concentration or temperature fields. As the marked fluid flows
through the light sheet, the dye is excited and re-emits light, so that the dye
concentration is
S
C (9)
E
This process is repeated for every interrogation zone in the whole region during a period
of time, in such a way that the dye concentration can be correlated with time and the
mixing time can be deducted. This technique allows simultaneously analyzing flow
structures in the tank [19, 66-72]. Distelhoff et al. [67] reported for the first time the use
of pLIF to determine mixing time in a baffled vessel containing water as working fluid.
In this case, the mixing time was defined as the time required for the dye to be
transported from the top of the vessel to the bottom plus the time required for a further
two rotations of the bulk flow in the circumferential direction. The mixing time was
based on the 99%-concentration because large variations were found when using lower
concentrations criteria. Arratia and Muzzio [73] reported the use of the pLIF technique
to measure the dye concentration in a laminar three-dimensional flow in a stirred vessel
equipped with three coaxial impellers. Figure 5 shows an example of the concentration
fields observed after 90 s, in which the effect of the Reynolds number on the mixing
time is clearly observed. They stated that the tracer variability decayed exponentially, so

12
ACCEPTED MANUSCRIPT

that the mixing time can be obtained once the concentration in the vessel reaches a
plateau with a minimum variability.

Figure 5 Concentration fields observed at 90 s: (a) Re = 40; (b) Re = 60. Concentration


scale bar in g/L [73].

Hu et al. [74] developed a technique to quantify the reactive process in an unbaffled

MANUSC
stirred tank using a novel reactive planar laser-induced fluorescence, which consists of
visualizing two liquids mixed and reacted with each other. For that purpose the

RIPT
fluorescence signal of the dye, which varies due to the presence of a reacting material
(H2O2), was continuously recorded along the reaction process. Figure 6 shows an
example of the concentration evolution, in which the red color corresponds to a tracer
concentration of 100 g/L, while the blue color denotes 0 g/L.

Figure 6 Normalized concentration fields [74].

13
ACCEPTED MANUSCRIPT

PLIF has been also used to characterize reactive and non-reactive mixing processes
in terms of time 99 and times 95 and 99, respectively [75]. They reported a time ratio
99 =99/99 (99% mixing time to 99% non-reactive mixing time) ranging from 0.42 to
0.58, which indicates that reactive mixing is mainly controlled by the corresponding
non-reactive mixing. As in colorimetry, simple digital image processing has been used
with concentration maps obtained with pLIF for determining mixing times [68,75,76].
The advantages and disadvantages as well as the operating principle of the planar
laser induced fluorescence are summarized in Table 4.

Table 4 Operating principle, advantage and disadvantage of Planar Laser Induced


Fluorescence technique (pLIF).
Operating principle Advantage Disadvantage
The flow region to be  Direct measurement  Calibration is required
analyzed is illuminated  Use in reacting and non- prior to the experiments
with a laser sheet and a reacting flows  Expensive technique
fluorescent tracer is fed.  Non-intrusive (cameras and laser)
Such tracer absorbs the  Instantaneous visualization  For transparent systems
laser light energy and re- with high temporal and  For semi-opaque fluids
emits light at a longer spatial resolution when using high power
wavelength that can be  Different flow field laser
detected by a variables can be obtained  Not suitable under
photodetector. (concentration, density, highly aerated
temperature) conditions
 Allows visualizing 2-D
isolated mixing regions
 For unaerated and low
aerated conditions

2.4 Thermography
Liquid crystal is an organic material in the amorphous solid form at a certain
temperature and pure liquid beyond the upper limit. Due to its molecular structure it
behaves as a crystal between these two phases. When an incident light is selectively

14
ACCEPTED MANUSCRIPT

scattered liquid crystals are the basis for temperature measurements. Based on the
thermography principle, Lee et al. [78] and Lee and Yianneskis [79] developed a liquid
crystal thermographic technique for the measurement of mixing time in a stirred vessel.
For that purpose, thermochromic liquid crystals encapsulated as gelatin-shell micro-
spheres having a mean diameter of 20 m were injected at the top surface. Since the
tracer and the working fluid had almost the same density the tracer remained suspended.
Figure 7 shows the hue contours of one-half of the flow field in the vessel stirred by 2-
Rushton turbine array rotating at 540 rpm, 200 ms after the tracer insertion. The tracer
was injected at z/T = 1.2 at a higher temperature than the working fluid, and the latter is
hued by red color.

CCEPT
ED

Figure 7 Hue contours of the flow field observed with two-Rushton turbines 200 ms
after tracer injection [79].

15
ACCEPTED MANUSCRIPT

The operating principle, advantages and disadvantages of the thermography


technique are summarized in Table 5.

Table 5 Operating principle, advantages and disadvantages of thermography


Operating principle Advantage Disadvantage
It is based on the change  Liquid crystals exhibit a  Use of sophisticated
in color of rapid and reversible image processing
thermochromic liquid response to dynamic software
crystals when they are temperature changes  Data processing is time
subjected to different  Wide range of consuming
temperatures. temperatures  Used in transparent
 Affordable setup liquids and transparent
 Non-intrusive tanks
 Temporal and spatial  Invasive technique
resolution  Mixing time must be
inferred from Eulerian
data
 Limited applicability
under highly aerated
conditions

2.5 Conductometry and pH


Physical measurements of conductivity and pH are made using probes placed at
different points into the stirred vessel. Irrespective of their size, the presence of such
probes perturb the flow, so that the measurement of mixing times should be considered
as apparent, especially when using very small vessels. The conductivity technique is
based on the measurement of the electrical conductivity of a solution over the time. For
that purpose a conductivity probe is placed into the stirred vessel at a specific position.
Figure 8a shows a typical setup of a conductivity measurement system, which consists
basically of a conductivity probe and a conductivity meter connected to a personal
computer through a data acquisition card.

16
ACCEPTED MANUSCRIPT

(a)

(b)
Figure 8 (a) Stirred vessel equipped with a conductivity measurement system; (b) Tracer
concentration fluctuating with time [3].

Basically, there two types of conductivity probes or cells, namely: 2-electrode cells
and 4-electrode cells. The first one is the most common conductivity probe made of
glass with two electrodes made of platinum. However, this kind of probes is prone to
polarization errors due to the electro-chemical reactions. To overcome this drawback,
the 4-electrode probe is the best option. In such a design, alternating current is applied
to the outer pair of rings and the voltage is measured on the inner rings avoiding
polarization effects since there are no current flows in the measuring circuit. A typical
conductivity probe is composed by anode and cathode, being the latter made from an
inert metal. The probe is placed into the solution and it is activated when voltage is
supplied. Then electron-carrying (negative) ions move towards the anode while
electron-less (positive) ions move towards the cathode. It is important to point that

17
ACCEPTED MANUSCRIPT

conductivity solutions are sensitive to temperature changes, therefore the system must
be temperature compensated or calibrated at the testing temperature.
The conductivity values are then converted into concentration data versus time
scale by using the calibration factor of the conductivity probe. In such a case, the initial
concentration should be measured and the mixing time requires the final concentration,
where the following expression must be small enough:
C(t)  C(0)
Ci  (11)
C()  C(0)
where C(0) is the initial tracer concentration, C() is the final tracer concentration and
C(t) is tracer concentration at a certain time. Figure 8b shows a typical plot of
concentration fluctuations of tracer as a function of time. As stated by Tatterson [3], the
amplitude of the concentration fluctuations decays exponentially with e Kat , where Ka is
a constant determined for various impellers and geometries. The data can be replotted in
terms of probe log variance as a function of time [78]. Under these conditions, the
mixing time can be determined once the concentration fluctuations are smaller than 5%,
which is known as the 95% mixing time (95).
Holmes et al. [81] reported for the first time the use of the conductivity technique to
perform a study of mixing effectiveness in a baffled stirred vessel. Using a conductivity
cell around the impeller, the average time required for a fluid element to complete one
circulation around the tank, which can be used for estimating the mixing time.
Bouwmans et al. [82] found that the probe position does not have any effect on the
mixing time measurement in the turbulent regime. Although the conductivity technique
has been used as an intrusive technique, Giona et al. [83] reported the use of impedance
probes attached to the baffles and another one mounted on the shaft, so that the flow in
the tank was not perturbed.
Most of the studies reporting the effect of different parameters on the measurement
of mixing time coincide that the mixing time in the laminar and turbulent regime is
strongly dependent on the following parameters [84-98]:
 D NQ V  
N  f  , , a ,  (13)
T Np V  
 m 

 
where D is the diameter of the impeller, T is the tank diameter, NQ/Np is pumping
effectiveness of the impeller, Va is the volume of the tracer added, V is the volume of the
liquid in the vessel and  is the density.

18
ACCEPTED MANUSCRIPT

The use of the conductivity technique has been also used for determining mixing
times in several applications such as gas dispersion [92,99], steel converters [100],
quality of mixing on animal cell culture [71], blending of fluids having different
densities [101], gassed conditions [102,105], etc.
On the other hand, Nagata [6] described a method based on local measurements of
pH for the determination of mixing times in stirred vessels. Following that approach,
Poulsen and Iversen [103] developed a system based on the signal of two pH sensors
placed in the bottom and top a stirred reactor and a conventional stimulus-response
technique, in which the tracer signal was the difference between two pH measurements,
so that the effect of the position of the probes in the tank was minimized and the initial
and final value of the tracer became zero being easier the determination of mixing
times. Figure 9 compares the determination of mixing time based on the measurement
of pH at the top (pH1) and at the bottom (pH2) of the vessel and the difference between
the two pH probes measurements (pH). As mixing time is slightly shorter when using
the difference of the two signals, the position of each probe in that case does not play an
important role.

ACCEP
TED

Figure 9 pH signal as a function of time: Signals from each probe (upper plot);
difference between the two probes (lower plot) (adapted from [103]).

19
ACCEPTED MANUSCRIPT

Vallejos et al. [104] developed a confocal optical to study mixing times in a stirred
reactor, whose results were validated by measuring pH at a specific position. Van der
Gulik et al. [106] reported also the use of the pulse-response technique for the
measurement of mixing time in a horizontal stirred reactor. For that purpose, they
perform the measurements outside the tank by circulating the reactor contents through a
spectrophotometer placed in an external loop.
The advantages, disadvantages and the operating principle of the technique based
on the use of probes (conductivity and pH) are summarized in Table 6.

Table 6 Operating principle, advantages and disadvantages of probes


Operating principle Advantage Disadvantage
Physical measurement of  Affordable setup  Invasive and intrusive
conductivity, pH, etc.  Simple to implement technique
using probes placed at  Direct measurements  Measurements of
different positions in the  For transparent and mixing times depend on
stirred vessel. opaque liquids the position of probes
 Suitable for unaerated and  Require calibration
aerated conditions  Measurement system to
be compensate for
temperature changes
 Limited temporal and
spatial resolution

3 EMPIRICAL CORRELATIONS
From the findings reported previously, mixing time can be predicted as a function
of the operating conditions. The mixing time correlations in Table 7 have been selected
from mixing studies using different measurement techniques as well as different
conditions, namely: flow regime, type of impeller, baffle and fluid rheology. From such
a table it is observed that the mixing time under the laminar regime correlates well with
the hydrodynamic characteristics, mainly on the agitation regime or the power drawn by
the impeller, which is true when mixing viscous fluids with close clearance impellers
such as helical ribbon impellers. However, in the turbulent regime under ungassed
conditions, mixing times is not only dependent on the agitation regime and the power

20
ACCEPTED MANUSCRIPT

consumption ( Np  2 NTm  N 3D5 ), but also on the tank geometry. Looking at the
ratio of impeller size (D) to tank size (T), it is observed than bigger impellers are more
effective than small impellers in the turbulent regime. However, in the same flow
regime but under gassed conditions, mixing time correlates with a number of parameters,
additionally to the latter case, the gas flow rate and the volume of liquid, among other
parameters. In this case, the mixing time is strongly dependent on the ratio of the
impeller size to the tank size being the exponent of the order of 2. On the other hand, the
gas flow rate does not play a significant role on the mixing time.

21
ACCEPTED MANUSCRIPT

Table 7 Mixing time correlations for stirred vessels


Laminar regime (ungassed)
Measurement
Correlation Geometry Fluid Ref.
technique
Helical ribbon Viscous [107] Conductometry
 m  896 x10 3 Kp 1.69

1  30 N R
/N
G
2 / 3  Triaxe® agitator
Newtonian
Viscous [36] Colorimetry

m  

Newtonian
NG 
Turbulent regime (ungassed)
2 Single impeller non- [10] Conductometry
N   5.2 Np 1/ 3D  
95  H/T = 1 Newtonian
T  baffled tank shear-thinning
 Single axial or radial Low viscosity [56] Electrical
2 b
1/ 3 D   H 
N  95  5.2 Np     impeller Newtonian resistance
 T   T  H/T = 2 tomography
Baffled
Multiple axial and radial Low viscosity [108] Conductometry
J 1
95  1.20 Ei  d 1 / 3 H 2 / 3 2 / 3 impellers Newtonian
J H/T > 1
unbaffled tank
Turbulent regime (gassed)
Nm  20.41  aH  T  1/12  1/15 Different type of impeller Low viscosity [102] pH and
 T   W   Q s   N 2 D4 
13/ 6

unbaffled tank Newtonian conductometry


 T  D   D  NV   gWV 
2 /3

2.43 Four coaxial radial or Low viscosity [105] Conductometry
H
 m  3.3 Np 1 / 3
N  
1
axial impellers Newtonian
D
G


95  12.5Q0.34 H 0.56 R1.93 Gas nozzles placed at the
bottom of the vessel
Low viscosity
Newtonian
[92] Conductometry

 m  A1  exp BRe Blade gas-entrainment Low viscosity [60] Electrical


Received date: Accepted date:
ACCEPTED MANUSCRIPT

impellers Newtonian resistance


tomography

23
ACCEPTED MANUSCRIPT

4 COMPARISON BETWEEN TECHNIQUES


The selection of a specific technique for the measurement of mixing time in stirred
vessels is dependent of several factors such as the following:
 Accuracy and reproducibility
 Type of information generated
 Suitability
 Costs
 Assembly of the experimental setup
 Calibration
 Sampling speed and data processing time
In general terms, the accuracy provided by techniques involving the use of tracers is
higher compared to physical methods. Colorimetry is by far the simplest technique in
which a color tracer is only needed; however, its accuracy and reproducibility is
strongly dependent on the subjectivity of human observer. This drawback can be
overcome by combining the technique with image processing methods. In that case,
higher resolution images will provide better results, but the cost will increase and the
data processing time could be longer. On the other hand, sophisticated techniques such
as electrical resistance tomography (ERT) are very accurate methods for the
measurement of mixing time. In such a case, the initial investment for acquiring those
systems should be considered, as well as the time for assembling the experimental setup
and the data processing time.
In regard to the type of information generated, it is important to point out that
experimental techniques providing Eulerian or Lagrangian data do not give a direct
measurement of mixing time, so the information obtained from these techniques must
be related to the mixing efficiency.
Other techniques, such as positron emission particle tracking (PEPT), computer
automated radioactive particle tracking (CARPT), particle image velocimetry (PIV),
laser Doppler anemometry (LDA) have been developed for visualization purposes.
However, mixing times can be indirectly inferred by following a sophisticated protocol.
For instance, García-Cortés et al. [109] reported the use of LDA data, from which
Eulerian information is generated and the mixing time can be deducted by means of a
mass balance as follows:

Received date: Accepted date:


ACCEPTED MANUSCRIPT

1. Radial profiles of the mean velocity ( vr ) in the vicinity of the impeller were
obtained.
2. The pumping capacity (Qp) was calculated by integrating the mean radial
velocity along the blade height times the cross sectional area.
3. The renewal time (tR) was determined as the ratio of the pumping capacity to the
volume of liquid, which is equivalent to the circulation time (m).
4. The mixing time was found to be proportional to the renewal time:
V
  m L  mt (14)
m R
Qp

where m is the number of time the discharge liquid circulates in the vessel,
which is a function of the tank diameter, the impeller geometry and the impeller
blade number.
On the other hand, the Lagrangian approach follows a fluid particle over time as it
moves through the flow field, then the mixing time can be determined from the flow
number or the circulation time. Nienow [110] found that macromixing mixing time
correlates well with the flow number, which is defined as the pumping capacity (Qp)
divided by ND3, being N the impeller rotational speed and D the impeller diameter. The
circulation time (c) is here defined as the ratio of volume of liquid (V) to the pumping
capacity.
In regard to the suitability, most of the techniques, especially those based on flow
visualization (e.g. colorimetry, pLIF) allow determining not only mixing times but also
obtaining 2-D and 3-D flow patterns in transparent vessels and fluids, which is a
limitation at industrial scales. This drawback can be alleviated using other techniques
like ERT, which can be applied to opaque systems. On the other hand, the use of probes
can provide also information about the hydrodynamics with opaque fluids. Techniques
involving the use of probes for the measurement of physical properties are in general
affordable systems and easy to be implemented; however, since the measurements are
local, a map showing the distribution of physical values could result in a time
consuming and costly process. From a physical standpoint, the use of probes can disturb
the flow, being this effect stronger when using small tanks.
Both calibration and data processing time are also important issues to consider
when selecting a measurement technique for mixing times. Particularly, pLIF is a
technique requiring calibration before use and sophisticated software is required for data

25
ACCEPTED MANUSCRIPT

processing. However, such a technique provides reliable information about flow


patterns, which can be further used for the accurate estimation of mixing time. Besides
calibration, some measurement techniques are sensitive to temperature; therefore they
need to be compensated for these changes (e.g. pLIF).
On the other hand, techniques based on probes provide almost real time
measurements. For instance, the measurements of conductivity or pH can be easily
correlated with mixing time in a stirred vessel if the probes are connected to a data
acquisition system and then such data is processed to determine the mixing time.
However, response time should be also considered, especially when using probes for the
measurement of conductivity or pH.
Mixing time has been not only determined by using experimental techniques but
also by numerical simulation [111-113]. All the techniques here described are useful
methods also for validating mixing time predicted with numerical simulations, such as
the one reported by Coroneo et al. [111], who performed simulations of fluid mixing in
a single phase of a stirred vessel with different levels of spatial discretization by finding
good agreement when comparing their results with pLIF results.
To summarize, besides the investment costs, the ideal experimental technique
would be the one which does not disturb the flow, it can be applied over a temperature
range, it can be used with opaque and transparent fluids and nontransparent tanks, and it
can provide accurate and reproducible results.

5 CONCLUSIONS
The experimental techniques developed in the last 50 years for measuring mixing
times in stirred vessels have been reviewed and the most outstanding works reporting
these techniques have been described. The applicability and limitations of all techniques
as well as a comparison between them in terms of the accuracy, type of information
generated, cost and processing time, among others, have been also described.
Techniques such as colorimetry and pLIF have been supplemented in the last years with
digital image processing resulting in more robust and accurate measurement methods.
On the other hand, relatively new techniques based on radioactive tracking like positron
emission particle tracking have emerged to provide information not only on the
homogeneity level in the tank, but also on the flow fields in non-transparent systems.

26
ACCEPTED MANUSCRIPT

NOMENCLATURE
Ac absorption phenomena factor (nondimensional)
b equation constant (Table 7) (nondimensional)
cT temperature compensation slope of the solution (nondimensional)
C concentration (nondimensional)
D impeller diameter (m)
E laser light intensity

impeller efficiency (   N Q3 / Np   D / T  , in Table 7)


4
Ei

foptic optical factor (nondimensional)


H liquid height (m)
J number of impellers (Table 7) (nondimensional)
k consistency index (power-law model) (nondimensional)
ks Metzner-Otto constant (nondimensional)
K mixing time constant (nondimensional)
Ka constant dependent on impeller type
L light path (nondimensional)
n flow behavior index (power-law model) (nondimensional)
N impeller rotational speed (1/s)
NG gyrational speed (1/s)
Np Power number (= Tm /  N 3D5  )
NpG gassed power number ( Tm N D3 ) 5

NQ flow or pumping number ( Q ND3 )

NQs secondary flow number ( Q ND3 )


NR rotational speed (1/s)
Q gas flow rate (m3)
Q quantum efficiency of the dye (nondimensional)
Qs sparged gas flow rate (m3)
R radius of vessel (m)
Re Reynolds number (  ND2  )
S concentration signal (nondimensional)
t time (s)
T tank diameter (m)

27
ACCEPTED MANUSCRIPT

Tcal calibration temperature (C)


Tliq temperature of liquid (C)
Tm torque (Nm)
vn normal velocity (m/s)
V volume of liquid (Table 7) (m3)
Va volume of tracer added (m3)
Vc sampling volume (m3)

vT mean velocity (m/s)


W mean tangential velocity (Table 7) (m/s)
z axial position (m)

Greek letters
 shear rate (1/s)
 spacing of interference light fringes (m)
t time difference (s)
X particle displacement (m)
 absorption index (nondimensional)
d mean energy dissipation rate (W/s)
95 95% time ratio (nondimensional)
99 99% time ratio (nondimensional)
 dynamic viscosity (Pas)
 density (kg/m3)
 conductivity (S/m)
90 90% mixing time (s)
95 95% mixing time (s)
99 99% mixing time (s)
c circulation time (s)
m mixing time (nondimensional)
 variance (nondimensional)
T c conductivity at temperature Tc (S/m)
T cal
conductivity at calibration temperature (S/m)

95 95% non-reactive mixing time (s)

28
ACCEPTED MANUSCRIPT

99 99% non-reactive mixing time (s)


 angular speed (rad/s)

REFERENCES
1. Harnby, N., Edwards, M.F., Nienow, A.W., Mixing in the Process Industries, Butterworth-
Heinemann, Oxford, UK (1997).
2. Patterson, G.K., Paul, E.L., Kresta, S.M., Etchells III, A.W., “Mixing and chemical reactions”
In: Paul, E.M., Kresta, S.M., Atiemo-Obeng, V.A., eds. Handbook of Industrial Mixing:
Science and Practice, John Wiley & Sons, New Jersey (2004)
3. Tatterson, G.B., Fluid Mixing and Gas Dispersion in Agitated Tanks, McGraw-Hill, USA
(1991).
4. Moo-Young, M., Tichar, K., Dullien, F.A.L., “The blending efficiencies of some impellers in
batch mixing”, AIChE J., 18, 178-182 (1972).
5. Metzner, A.B., Otto, R.E., “Agitation of non-Newtonian fluids”, AIChE J., 3, 3-10 (1957)
6. Nagata S. Mixing: Principles and Applications, Wiley, New York (1975).
7. Nishikawa, M., Ashiwake, K., Hashimoto, N., Nagata, S., “Agitation power and mixing time in
off-centering mixing”, Int. Chem. Eng., 19, 153-159 (1979).
8. Nere, N.K., Patwardhan, A.W., Joshi, J.B., “Liquid-phase mixing in stirred vessels: Turbulent
flow regime”, Ind. Eng. Chem. Res., 42, 2661-2698 (2003).
9. Sano, Y., Usui, H., “Effects of paddle dimensions and baffle conditions on the interrelations
among discharge flow rate, mixing power and mixing time in mixing vessels”, J. Chem. Eng.
Jpn., 20, 399-404 (1987).
10. Grenville, R.K., Nienow, A.W., “Blending of miscible liquids”, In: Paul, E.M., Kresta, S.M.,
Atiemo-Obeng, V.A., editors. Handbook of Industrial Mixing: Science and Practice, 1 st ed.,
John Wiley, New Jersey (2004).
11. Szoplik J, Karcz J. An efficiency of the liquid homogenization in agitated vessels equipped
with off-centred impellers. Chem. Pap., 59, 373-379 (2005).
12. Mavros, P., “Flow visualization in stirred vessels. A review of experimental techniques”, Chem.
Eng. Res. Des., 79, 113-127 (2001).
13. Hoogendoorn, C.J., den Hartog, A.P., “Model studies on mixers in the viscous flow region”,
Chem. Eng. Sci., 22, 1689-1699 (1967).
14. Carreau, P.J., Patterson, I., Yap, C.Y., “Mixing of viscoelastic fluids with helical-ribbon
agitators. I - Mixing time and flow patterns”, Can. J. Chem. Eng., 54, 135-142 (1976).
15. Kraume, M., Zehner, P., “Experience with experimental standards for measurements of various
parameters in stirred tanks: A comparative test”, Chem. Eng. Res. Des., 79, 811-818 (2001).
16. Lamberto, D.J., Muzzio, F.J., Swanson, P.D., Tonkovich, A.L., “Using time-dependent RPM to
enhance mixing in stirred vessels”, Chem. Eng. Sci., 51, 733-741 (1996).

29
ACCEPTED MANUSCRIPT

17. Yao, W.G., Sato, H., Takahashi, K., Koyama, K., “Mixing performance experiments in impeller
stirred tanks subjected to unsteady rotational speeds”, Chem. Eng. Sci., 53, 3031-3040 (1998).
18. Ascanio, G., Brito-Bazán, M., Brito-de la Fuente, E., Carreau, P.J., Tanguy, P.A.,
“Unconventional configuration studies to improve mixing times in stirred tanks”, Can. J. Chem.
Eng., 80, 558-565 (2002).
19. Alvarez, M., Arratia, P.E., Muzzio, F.J., “Laminar mixing in eccentric stirred tank systems”,
Can. J. Chem. Eng., 80, 546-557 (2002).
20. Ascanio, G., Foucault, S., Tanguy, P.A., “Time-periodic mixing of shear-thinning fluids”,
Chem. Eng. Res. Des., 82, 1199-1203 (2004).
21. Hidalgo-Millán A., “Geometric perturbations in mechanically agitated tanks”, Ph.D.
dissertation, National Autonomous University of Mexico (2010). (in Spanish)
22. Norwood, K.W., Metzner, A.B., “Flow patterns and mixing rates in agitated vessels”, AIChE J.,
6, 432-437 (1960).
23. Hari-Prajitno, D., Mishra, V.P., Takenaka, K., Bujalski, W., Nienow, A.W., Mckemmie, J.,
“Gas-liquid mixing studies with multiple up- and down-pumping hydrofoil impellers: Power
characteristics and mixing time”, Can. J. Chem. Eng., 76, 1056–1068 (1998).
24. Pandit, A.B., Rielly, C.D., Niranjan, K., Davidson, J.F., “The convex bladed mixed flow
impeller and the marine propeller: A multipurpose agitator”, Chem. Eng. Sci., 44, 2463-2474
(1989).
25. Wang, S.-J., Zhong, J.-J., “A novel centrifugal impeller bioreactor. I. Fluid circulation, mixing,
and liquid velocity profiles”, Biotech. Bioeng., 51, 511-519 (1996).
26. Espinosa-Solares, T., Brito-de la Fuente, E., Tecante, A., Medina-Torres, L., Tanguy, P.A.,
“Mixing time in rheologically evolving model fluids by hybrid dual mixing systems”, Chem.
Eng. Res. Des., 80, 817-823 (2001).
27. Foucault, S., Ascanio, G., Tanguy, P.A., “Mixing Times in Coaxial Mixers with Newtonian and
non-Newtonian Fluids”, Ind. Eng. Chem. Res., 45, 352-359 (2006).
28. Hirata Y, Ito R. Characteristics of flow and mixing in vessel with rotating multistage disks, Ing:
Proceedings of 6th European Conference on Mixing, Pavia, Italy, (1988).
29. Hiruta, O., Yamamura, K., Takebe, H., Futamura, T., Iinuma, K., Tanaka, H., “Application of
Maxblend fermentor® for microbial processes”, J. Ferment. Bioeng., 83, 79-86 (1997).
30. Iranshahi, A., Devals, C., Heniche, M., Fradette, L., Tanguy, P.A., Takenaka, K.,
“Hydrodynamics characterization of the Maxblend impeller”, Chem. Eng. Sci., 62, 3641-3653
(2007).
31. Takahashi, K., Yokota, T., Furukawa, T., Harada, K., “Mixing of highly viscous Newtonian
liquid in a helical ribbon agitated vessel at various liquid depths”, J. Chem. Eng. Jpn., 27, 244-
247 (1994).

30
ACCEPTED MANUSCRIPT

32. Takahashi, T., Tagawa, A., Atsumi, N., Dohi, N., Kawase, Y., “Liquid-phase mixing time in
boiling stirred tank reactors with large cross-section impellers”, Chem. Eng. Process., 45, 303-
311 (2006).
33. Takahashi, K., Sugo, Y., Takahata, Y., Sekine, H., Nakamura, M., “Laminar mixing in stirred
tank agitated by an impeller inclined”, Int. J. Chem. Eng., article ID 858329 (2012).
34. Aizawa, E., Sakano, N., Imakoma, H., Ohmura, N., “Effect of rheological property of fluids on
mixing time in a stirred vessel”, Kagaku Kogaku Ronbun, 35, 539-542 (2009).
35. Kouda, T., Yano, H., Yoshinaga, F., Kaminoyama, M., Kamiwano, M., “Characterization of
non-Newtonian behavior during mixing of bacterial cellulose in a bioreactor”, J. Ferment.
Bioeng., 82, 382-386 (1996).
36. Delaplace, G., Bouvier, L., Moreau, A., Guérin, R., Leuliet, J.C., “Determination of mixing
time by colourimetric diagnosis - Application to a new mixing system”, Exp. Fluids 36, 437-
443 (2004).
37. Chhabra, R.P., Bouvier, L., Delaplace, G., Cuvelier, G., Domenek, S., André, C.,
“Determination of mixing times with helical ribbon impeller for non-Newtonian viscous fluids
using an advanced imaging method”, Chem. Eng. Technol., 30, 1686-1691 (2007).
38. Visuri, O., Laakkonen, M., Aittamaa, J., “A digital imaging technique for the analysis of local
inhomogeneities from agitated vessels”, Chem. Eng. Tech., 30, 1692-1699 (2007).
39. Cabaret, F., Bonnot, S., Fradette, L., Tanguy, P.A., “Mixing time analysis using colorimetric
methods and image processing”, Ind. Eng. Chem. Res., 46, 5032-5042 (2007).
40. Vega-Alvarado, L., Taboada, B., Hidalgo-Millán, A., Ascanio, G., “Image Analysis Method for
the Measurement of Mixing Times in Stirred Vessels”, Chem. Eng. Tech., 34, 859-866 (2011).
41. ITS-Industrial Tomography Systems Plc: http://www.itoms.com (visited on February 2013).
42. Tapp, H.S., Williams, R.A., “Status and applications of microelectrical resistance tomography”,
Chem. Eng. J., 77, 119-125 (2000).
43. Holden, P.J., Wang, M., Mann, R., Dickin, F.J., Edwards, R.B., “Imaging stirred-vessel
macromixing using electrical resistance tomography”, AIChE J., 44, 780-790 (1998).
44. Mann, R., Dickin, F.J., Wang, M., Dyakowski, T., Williams, R.A., Edwards, R.B., Forrest,
A.E., Holden, P.J., “Application of electrical resistance tomography to interrogate mixing
processes at plant scale”, Chem. Eng. Sci., 52, 2087-2097 (1997).
45. Mann, R., Wang, M., Forrest, A.E., Holden, P.J., Dyakwski, F.J., “Gas-liquid and miscible
liquid mixing in a plant-scale vessel monitored using electrical resistance tomography”, Chem.
Eng. Comm., 175, 33-48 (1999).
46. Holden, P.J., Wang, M., Mann, R., Dickin, F.J., Edwards, R.B., “On detecting mixing
pathologies inside a stirred vessel using electrical resistance tomography”, Chem. Eng. Res.
Des., 77, 709-712 (1999).
47. Pinheiro, P.A.T., Loh, W.W., Waterfall, R.C., Wang, M., Mann, R., “Three-dimensional
electrical resistance tomography in a stirred vessel”, Chem. Eng. Comm., 175, 25-38 (1999).

31
ACCEPTED MANUSCRIPT

48. 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., 77, 93-98 (2000).
49. Stanley, S.J., Mann, R., Primrose, K., “Tomographic imaging of fluid mixing in three
dimensions for single-feed semi-batch operation of a stirred vessel”, Chem. Eng. Res. Des., 80,
903-909 (2002).
50. Stanley, S.J., Mann, R., Primrose, K., “Interrogation of a precipitation reaction by electrical
resistance tomography (ERT)”, AIChE J., 51, 607–614 (2005).
51. Stanley, S.J., “Tomographic imaging during reactive precipitation in a stirred vessel: Mixing
with chemical reaction”, Chem. Eng. Sci., 61, 7850–7863 (2006).
52. Pakzad, L., Ein-Mozaffari, F., Chan, P., “Measuring mixing time in the agitation of non-
Newtonian fluids through electrical resistance tomography”, Chem. Eng. Technol., 31, 1838-
1845 (2008).
53. Rodgers, T.L., Gangolf, L., Vannier, C., Parriaud, M., Cooke, M., “Mixing times for process
vessels with aspect ratios greater than one”, Chem. Eng. Sci., 66, 2935-2944 (2011).
54. Rodgers, T.L., Kowalski, A., “An electrical resistance tomography method for determining
mixing in batch addition with a level change”, Chem. Eng. Res. Des., 88, 204–212 (2010).
55. Rodgers, T.L., Cooke, M., Siperstein, F.R., Kowalski, A., “Mixing and dissolution times for a
cowles disk agitator in large-scale emulsion preparation”, Ind. Eng. Chem. Res., 48, 6859-6868
(2009).
56. Rodgers, T.L., Siperstein, F.R., Mann, R., York, T.A., Kowalski, A., “Comparison of a
networks-of-zones fluid mixing model for a baffled stirred vessel with three-dimensional
electrical resistance tomography”, Meas. Sci. Technol., 22, article 104014 (2011).
57. Pakzad, L., Ein-Mozaffari, F., Upreti, S.R., Lohi, A., “Characterisation of the mixing of non-
Newtonian fluids with a Scaba 6SRGT impeller through ERT and CFD”, Can. J. Chem. Eng.,
91, 90-100 (2013).
58. Mann, R., “ERT imaging and linkage to CFD for stirred vessels in the chemical process
industry”, In: Proceedings of IEEE International Workshop Imaging Systems, IST 2009, Hong
Kong (2009).
59. Wang, M., Dorward, A., Vlaev, D., Mann, R., “Measurements of gas-liquid mixing in a stirred
vessel using electrical resistance tomography (ERT)”, In: Proceedings of the 1st World
Congress Industrial Process Tomography, Buxton, Greater Manchester, (1999).
60. Abdullah, B., Dave, C., Nguyen, T.-H., Cooper, C.G., Adesina, A.A., “Electrical resistance
tomography-assisted analysis of dispersed phase hold-up in a gas-inducing mechanically stirred
vessel”, Chem. Eng. Sci., 66, 5648–5662 (2011).
61. Dantec Dynamics, Inc.: http://www.dantecdynamics.com (visited on February 2014).
62. Edwards, I., Axon, S.A., Barigou, M., Stitt, E.H., “Combined use of PEPT and ERT in the
study of aluminum hydroxide precipitation”, Ind. Eng. Chem. Res., 48, 1019-1028 (2009).
63. PEPT Cape Town: http://www.pept.uct.ac.za/ (visited on February 2013)

32
ACCEPTED MANUSCRIPT

64. 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., 55, 5969-
5979 (2000).
65. 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., 42, 839-
846 (2009).
66. Yek, W.M., Noui-Mehidi, M.N., Parthasarathy, R., Bhattacharya, S.N., Wu, J., Ohmura, N.,
Nishioka, N., “Enhanced mixing of Newtonian fluids in a stirred vessel using impeller speed
modulation”, Can. J. Chem. Eng., 87, 839–846 (2009).
67. Distelhoff, M.F.W., Marquis, A.J., Nouri, J.M., Whitelaw, J.H., “Scalar mixing measurements
in batch operated stirred tanks”, Can. J. Chem. Eng., 75, 641-652 (1997).
68. Hall, J.F., Barigou, M., Simmons, M.J.H., Stitt, E.H., “Mixing in unbaffled high-throughput
experimentation reactors”, Ind. Eng. Chem. Res., 43, 4149-4158 (2004).
69. 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., 85, 560-567
(2007).
70. Zadghaffari, R., Moghaddas, J.S., Revstedt, J., “A mixing study in a double-Rushton stirred
tank”, Comp. Chem. Eng., 33, 1240-1246 (2009).
71. Collignon, M.L., Dossin, D., Delafosse, A., Crine, M., Toye, D., “Quality of mixing in a stirred
bioreactor for animal cells culture: heterogeneities in a lab scale bioreactor and evolution of
mixing time with scale up”, Biotechnol. Agron. Environ, 14(S2), 585-591 (2010).
72. Busciglio, A., Grisafi, F., Scargiali, F., Brucato, A., “Mixing time in unbaffled stirred tanks”,
In: Proceedings of 14th European Conference of Mixing, Warszawa, Poland (2012).
73. Arratia, P.E., Muzzio, F.J., “Planar laser-Induced fluorescence method for analysis of mixing in
laminar flows”, Ind. Eng. Chem. Res., 43, 6557-6568 (2004).
74. Hu, Y., Liu, Z., Yang, J., Jin, Y., Cheng, Y., “Study on the reactive mixing process in an
unbaffled stirred tank using planar laser-induced fluorescence (pLIF) technique”, Chem. Eng.
Sci., 65, 4511-4518 (2010).
75. Hu, Y., Wang, W., Shao, T., Yang, J., Cheng, Y., “Visualization of reactive and non-reactive
mixing processes in a stirred tank using planar induced fluorescence (pLIF) technique”, Chem.
Eng. Res. Des., 90, 524-533 (2012).
76. Houcine, I., Vivier, H., Plasari, E., David, R., Villermaux, J., “Planar laser induced
fluorescence technique for measurements of concentration fields in continuous stirred tank
reactors”, Exp. Fluids, 22, 95-102 (1996).
77. Fall, A., Lecoq, O., David, R., ”Characterization of mixing in a stirred tank by planar laser
induced fluorescence (P.L.I.F.)”, Chem. Eng. Res. Des., 79, 876–882 (2001).

33
ACCEPTED MANUSCRIPT

78. Lee, K.C., Yianneskis, M., Bertrand, J., Villermaux, J., “Measurement of temperature and
mixing time in stirred vessels with liquid crystal thermography”, In: Proceedings of 9th
European Conference on Mixing, Paris, France (1997).
79. Lee, K.C., Yianneskis, M., “A liquid crystal thermographic technique for the measurement of
mixing characteristics in stirred vessels”, Chem. Eng. Res. Des., 75, 746-754 (1997).
80. Brown, D.A.R., Jones, P.N., Middleton, J.C., “Experimental methods, Part A: Measuring tools
and techniques for mixing and flow visualization studies”, In: Paul, E.M., Kresta, S.M.,
Atiemo-Obeng, V.A., editors. Handbook of Industrial Mixing: Science and Practice, 1 st ed.,
John Wiley & Sons, New Jersey (2004)
81. Holmes, D.B., Voncken, R.M., Dekker, J.A., “Fluid flow in turbine-stirred, baffled tanks-I.
Circulation time”, Chem. Eng. Sci., 19, 201-208 (1964).
82. Bouwmans, I., Barker, A., van Den Akker, H.E.A., “Blending liquids of differing viscosities
and densities in stirred vessels”, Chem. Eng. Res. Des., 75, 777-783 (1997).
83. Giona, M., Paglianti, A., Cerbelli, S., Pintus, S., Adrover, A., “Tracer dispersion in stirred tank
reactors: Asymptotic properties and mixing characterization”, Can. J. Chem. Eng., 80, 580-590
(2002).
84. Sano, Y., Usui, H., “Interrelations among mixing time, power number and discharge flow rate
number baffled mixing vessels”, J. Chem. Eng. Jpn., 18, 47-52 (1985).
85. Jahoda, M., Machon, V., “Homogenization of liquids in tanks stirred by multiple impellers”,
Chem. Eng. Tech., 17, 95-101 (1994).
86. Rutherford, K., Lee, K.C., Mahmoudi, S.M.S., Yianneskis, M., “Hydrodynamic characteristics
of dual Rushton impeller stirred vessels”, AIChE J., 42, 332-346 (1996).
87. Becker, J.-U., Oeters, F., “Model experiments of mixing in steel ladles with continuous addition
of the substance to be mixed”, Steel Res. Int., 69, 8-16 (1998).
88. Gogate, P.R., Pandit, A.B., “Mixing of miscible liquids with density differences: Effect of
volume and density of the tracer fluid”, Can. J. Chem. Eng., 77, 988-996 (1999).
89. Kamei, N., Hiraoka, S., Kato, Y., Tada, Y., Yamazaki, K., “Effects of impeller and baffle
conditions on mixing time in turbulent agitated vessels”, Kagaku Kogaku Ronbun, 28, 9-15
(2002).
90. Michelett, M., Nikiforaki, L., Lee, K.C., Yianneskis, M., “Particle concentration and mixing
characteristics of moderate-to-dense solid-liquid suspensions”, Ind. Eng. Chem. Res., 42, 6236-
6249 (2003).
91. Kasat, G.R., Pandit, A.B., “Mixing time studies in multiple impeller agitated reactors”, Can. J.
Chem. Eng., 82, 892-904 (2004).
92. Mandal, J., Patil, S., Madan, M., Mazumdar, D., “Mixing time and correlation for ladles stirred
with dual porous plugs”, Metall. Mater. Trans. B, 36, 479-487 (2005).
93. Delvigne, F., Destain, J., Thonart, P., “Structured mixing model for stirred bioreactors: An
extension to the stochastic approach”, Chem. Eng. J., 113, 1-12 (2005).

34
ACCEPTED MANUSCRIPT

94. Cents, A.H.G., Jansen, D.J.W., Brilman, D.W.F., Versteeg, G.F., “Influence of small amounts
of additives on gas hold-up, bubble size, and interfacial area”, Ind. Eng. Chem. Res., 44, 4863-
4870 (2005).
95. Kumaresan, T., Nere, N.K., Joshi, J.B., “Effect of internals on the flow pattern and mixing in
stirred tanks”, Ind. Eng. Chem. Res., 44, 9951-9961 (2005).
96. Hasal, P., Jahoda, M., Fořt, I., “Free liquid surface motions in a stirred tank: An insight into the
fluid flow dynamics”, In: Proceedings of CHISA 2006 - 17th International Congress Chemical
and Process Engineering, Prague - Czech Republic (2006).
97. Buwa, V., Dewan, A., Nassar, A.F., Durst, F., “Fluid dynamics and mixing of single-phase flow
in a stirred vessel with a grid disc impeller: Experimental and numerical investigations”, Chem.
Eng. Sci., 61, 2815-2822 (2006).
98. Bao, Y., Chen, L., Gao, Z., Chen, J., “Local void fraction and bubble size distributions in cold-
gassed and hot-sparged stirred reactors”, Chem. Eng. Sci., 65, 976-984 (2010).
99. Otomo, N., Bujalski, W., Nienow, A.W., Takahashi, K., “A novel measurement technique for
mixing time in an aerated stirred vessel”, J. Chem. Eng. Jpn., 36, 66-74 (2003).
100. Martín, M., Rendueles, M., Díaz, M., “Global and local mixing determinations for steel
converter analysis”, Chem. Eng. Sci., 60, 5781-5791 (2005).
101. Jones, P.N., Özcan-Taşkin, G.N., “Effects of physical property differences on blending”, Chem.
Eng. Tech., 28, 908-914 (2005).
102. Pandit, A.B., Joshi, J.B., “Mixing in mechanically agitated gas-liquid contactors, bubble
columns and modified bubble columns”, Chem. Eng. Sci., 38, 1189-1215 (1983).
103. Poulsen, B.R., Iversen, J.J.L., “Mixing determinations in reactor vessels using linear buffers”,
Chem. Eng. Sci., 52, 979-984 (1997).
104. Vallejos, J.R., Kostov, Y., Marten, M.R., Rao, G., “Confocal optical system: A novel non-
invasive sensor to study mixing”, Biotechnol. Prog., 21, 1531-1536 (2005).
105. Vrábel, P., van der Lans, R.G.J.M., Luyben, K.Ch.A.M., Boon, L., Nienow, A.W., “Mixing in
large-scale vessels stirred with multiple radial or radial and axial up-pumping impellers:
modeling and measurements”, Chem. Eng. Sci., 55, 5881-5896 (2000).
106. van der Gulik, G.J.S., Wijers, J.G., Keurentjes, J.T.F., “Hydrodynamics and scale-up of
horizontal stirred reactors”, Ind. Eng. Chem. Res., 40, 4731-4740 (2001).
107. Grenville, R.K., Hutchinson, T.M., Higbee, R.W., “Optimisation of helical ribbon geometry
for blending in the laminar regime”, In: Proceedings of Mixing XVIII, North American Mixing
Forum, Pocono, U.S.A, (2001).
108. Magelli, F., Montante, G., Pinelli, D., Paglianti, A., “Mixing time in high aspect ratio vessels
stirred with multiple impellers”, Chem. Eng. Sci., 101, 712-720 (2013).
109. García-Cortés, D., Ferrer, E., Barberà, E., “Hydrodynamic characterization of the flow
induced by a four-bladed disk-style turbine”, Chem. Eng. Res. Des., 79, 269-273 (2001).

35
ACCEPTED MANUSCRIPT

110. Nienow, A.W., “On the impeller circulation and mixing effectiveness in the turbulent flow
regime”, Chem. Eng. Sci., 52, 2557–2565 (1997).
111. Coroneo, M., Montante, G., Paglianti, A., Magelli, F., “CFD prediction of fluid flow and
mixing in stirred tanks: Numerical issues about the RANS simulations”, Comp. Chem. Eng., 35,
1959-1968 (2011).
112. Min, J., Gao, Z., Shi, L., “CFD simulation of mixing in a stirred tank with multiple hydrofoil
impellers”, Chinese J. Chem. Eng., 13, 583-588 (2005).

113. Min, J., Gao, Z., “Large Eddy Simulations of Mixing Time in a stirred vessel”, Chinese J.
Chem. Eng., 14, 1-7 (2006).

36

You might also like