Download as pdf or txt
Download as pdf or txt
You are on page 1of 48

Subscriber access provided by University of South Dakota

Thermodynamics, Transport, and Fluid Mechanics


Influence of impeller geometry on hydromechanical
stress in stirred liquid/liquid dispersions
Chrysoula Bliatsiou, Alexander Malik, Lutz Böhm, and Matthias Kraume
Ind. Eng. Chem. Res., Just Accepted Manuscript • DOI: 10.1021/acs.iecr.8b03654 • Publication Date (Web): 18 Dec 2018
Downloaded from http://pubs.acs.org on December 19, 2018

Just Accepted

“Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted
online prior to technical editing, formatting for publication and author proofing. The American Chemical
Society provides “Just Accepted” as a service to the research community to expedite the dissemination
of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in
full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully
peer reviewed, but should not be considered the official version of record. They are citable by the
Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore,
the “Just Accepted” Web site may not include all articles that will be published in the journal. After
a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web
site and published as an ASAP article. Note that technical editing may introduce minor changes
to the manuscript text and/or graphics which could affect content, and all legal disclaimers and
ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or
consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W.,


Washington, DC 20036
Published by American Chemical Society. Copyright © American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the course
of their duties.
Page 1 of 47 Industrial & Engineering Chemistry Research

1
2
3
4
5
6
7
8
Influence of impeller geometry on hydromechanical
9
10
11
12
stress in stirred liquid/liquid dispersions
13
14
15
16 Chrysoula Bliatsiou*, Alexander Malik, Lutz Böhm, Matthias Kraume
17
18
19
20
Chair of Chemical and Process Engineering, Technische Universität Berlin, FH6-1, Straße des
21
22 17. Juni 135, 10623 Berlin, Germany
23
24
25 KEYWORDS. stirred tank, impeller, particle stress, multiphase flow, drop size distribution,
26 liquid/liquid dispersion
27
28
29 ABSTRACT. Hydromechanical stress is a crucial parameter for a broad range of multiphase
30
31 processes in the field of (bio-)chemical engineering. The effect of impeller type and geometry on
32
33
34
hydromechanical stress in stirred tanks is important. The present study aims at characterizing
35
36 conventional and new impeller types in terms of particle stress. A two-phase liquid/liquid non-
37
38 coalescing dispersion system is employed, and the drop breakage is monitored in-line in a stirred
39
40
tank. The published effects of agitation on drop deformation were confirmed and expanded
41
42
43 significantly for five modified new impeller types. Radial impellers are advantageous for
44
45 applications where low shear conditions are desired. A modified propeller with a peripheral ring
46
47 and the developed wave-ribbon impellers present remarkable results by producing significantly
48
49
50 low and high hydromechanical stress respectively. The results obtained were correlated in terms
51
52 of mean and maximum energy dissipation rate, as well as circulation frequency in the impeller
53
54 swept volume.
55
56
57
58
59
60 ACS Paragon Plus Environment
1
Industrial & Engineering Chemistry Research Page 2 of 47

1
2
3 INTRODUCTION
4
5
6
7 In industrial application, stirred tank reactors are broadly used in chemical and biochemical
8
9
processes. Stirring is a crucial process parameter, which must fulfill several and often conflicting
10
11
12 goals, i.e., bulk fluid mixing, multiphase dispersion, heat and mass transfer, homogeneity of the
13
14 processed material, shear stress on suspended particles. For systems with suspended particles
15
16 (e.g., crystals, flocs, drops), or growing organisms (e.g., biological cells, microorganisms) the
17
18
19 stirring conditions define the mechanical stress on the suspended material. The particle stress can
20
21 often be desired to produce an increased interfacial area between two phases in emulsification,
22
23 polymerization, dispersion, and aeration processes.1–6 On the other hand, processes such as
24
25
26
crystallization, precipitation, or biotechnological applications, where shear-sensitive
27
28 microorganisms or cell structures are present, can be detrimentally affected by particle stress. 7–11
29
30 Whether beneficial or not, the hydromechanical stress is of crucial significance for the design
31
32
and operation of chemical and biochemical processes.
33
34
35 Particle stress is mainly the result of the relative velocity between particles and the surrounding
36
37 fluid. Most frequently, dispersion processes in stirred tanks are performed under turbulent flow
38
39 conditions.12,13 The stress acting on particles is determined by the local turbulent velocity
40
41
2
42 fluctuations over time √u' . Additionally, if there is a significant density difference between the
43
44
two phases, i.e., bulk fluid and particles, the mean velocities of the particles and the fluid also
45
46
47 differ, which leads to impact stress.12,14 The latter is defined as the stress caused either by the
48
49 contact between particles, between a particle and the impeller element, or between a particle and
50
51 the tank wall. However, when the density differences or the particle concentrations are low, the
52
53
54 impact stress is considered negligible.12,14
55
56
57
58
59
60 ACS Paragon Plus Environment
2
Page 3 of 47 Industrial & Engineering Chemistry Research

1
2
3 In a stirred tank the turbulence of flow can be described by the theory of energy cascade.15
4
5
6 Under turbulent flow conditions eddies of different sizes are formed. The turbulent kinetic
7
8 energy is transferred successively from large energy-rich eddies to smaller ones with a decreased
9
10 energy content. This process is limited by the viscosity of the fluid, which causes the kinetic
11
12
13
energy to be ultimately dissipated as thermal energy in eddies of minimal size present in the
14
15 fluid.12,15 Under isotropic turbulence conditions, these terminal eddies have a characteristic
16
17 ¼
 3
18 length  known as the Kolmogorov microscale  = ( fl ⁄loc ) , where fl is the kinematic
19
20
21
viscosity of the fluid and loc the local specific energy dissipation rate.16 In stirred reactors,
22
23
24 although the main flow is not isotropic, local isotropy can be assumed.15 The maximum local
25
26 energy dissipation rate max occurs in the impeller region and defines the minimum size of eddies
27
28
that are formed in the stirred tank. For the hydromechanical particle stress, the size ratio between
29
30
31 particles and flow eddies is of crucial significance. Particle disintegration is mainly caused by
32
33 eddies with a size comparable to that of the particles.12,15 When the eddies are significantly larger
34
35 than the suspended particles, the particles follow them in a convective movement. Eddies much
36
37
38 smaller than the particles have an energy content which is too low to induce breakage. In the
39
40 velocity field of the determining eddies, the dynamic stress acting on particles is defined
41
42 ̅̅̅̅2 . Considering the distance r = dP between
according to the Reynolds stress t = ∙u´
43
44
45 neighboring points in the flow field, the stress is given by: 12,15
46
2
47
Inertial range: r > 25∙ , t ∝∙(loc ∙Δr)3 (1)
48
49 loc
50 Dissipation range: r < 6∙, t ∝∙Δr2 ∙ (2)
fl
51
52 1
53 
When r < 3∙ , the shear stress becomes independent of r, with t ∝∙( loc )2 .17
54 fl
55
56
57
58
59
60 ACS Paragon Plus Environment
3
Industrial & Engineering Chemistry Research Page 4 of 47

1
2
3 In stirred systems, the particle stress has been correlated with different process parameters,
4
5 P
6 such as the mass specific average energy dissipation rate ̅ = , the impeller tip speed utip , the
∙V
7
8
9 maximum local energy dissipation rate max , the circulation time tc , and the energy dissipation
10
11 circulation function (EDCF).
12
13 The mean specific energy dissipation rate and the impeller tip speed have been insufficient to
14
15
16 interpret the stress phenomena in stirred systems, where local characteristics of turbulence play a
17
18 significant role, and thus they have been proven to be impractical criteria for scale-up.7,18,19 On
19
20 the other hand, direct measurements of the local energy dissipation are experimentally difficult.20
21
22
23
According to published studies up to 60% or even 80% of the energy dissipation occurs in the
24
25 impeller region.21–23 In this region the most severe breakage of suspended particles is expected to
26
27 max
happen and as a result the ratio of maximum local to mass-averaged energy dissipation rate 
is
28
29
30 significant for the hydromechanical stress caused by a given impeller and reactor geometry.
31
32 max
Table 1 presents correlations available in literature to estimate as a function of the impeller
33 
34
35 and reactor geometrical characteristics. These correlations often lead to contradictory results for
36
37 the hydromechanical stress caused by different impeller types (axial, radial). Works based on
38
39
40
fluid dynamic measurements (e.g. Particle Image Velocimetry-PIV, Laser Doppler Anemometry-
41
42 LDA), such as the one published by Geisler24, report that axial impellers with low power
43
44 max
numbers are characterized by low 
ratios. For this reason, axial stirrers were considered to be
45
46
47 “low-shear” agitators in the past.25 Indirect methods to estimate the particle stress caused by
48
49 different stirrers were applied by other researchers. Henzler and Biedermann14 and Pacek et al.26
50
51
used two-phase particulate systems for their investigations (floc suspensions and drop
52
53
54 dispersions respectively) and they recorded the particle breakage that occurred during stirring.
55
56 These studies together with the work of Jüsten et al.18, who worked on the fragmentation of
57
58
59
60 ACS Paragon Plus Environment
4
Page 5 of 47 Industrial & Engineering Chemistry Research

1
2
3 filamentous bioagglomerates, were the first to conclude that axial stirrers cause higher
4
5
6 hydromechanical stress leading to an increased particle breakage. These results could indicate
7
max
8 higher ratios for the case of the axial impellers in comparison to the radial ones. The

9
10 max
11 contradictions among the various approaches to estimate are represented by the mathematical

12
13
models summarized in Table 1. McManamey estimated the maximum dissipation rate by
14
15
16 accepting that the main energy dissipation occurs in the impeller region.27 Instead of the total
17
18 tank volume, he used the impeller swept volume VI to evaluate the order-of-magnitude of the
19
20 maximum energy dissipation. This approach has been adopted widely in the literature to
21
22 max
23 correlate particle sizes with in mixing systems.26,28,29 Henzler and Biedermann proposed that

24
25
for the definition of the impeller volume VD , where the largest part of the energy dissipates, the
26
27
28 geometrical features of the impeller have to be considered.14 They introduced a geometrical
29
max
30 function F, which allows the approximation of . To interpret the particle stress in a universal
31 
32
33 way by using only the maximum to average energy dissipation rate is challenging.13,24 To
34
35 overcome this challenge, it has often been stated that apart from the maximum energy
36
37 dissipation, the local distribution of energy dissipation and the local residence time of particles in
38
39
40 the respective energy dissipation zones should be taken into consideration to assess the particle
41
42 stress in stirred systems.19, 24,30
43
44 In 1996 Jüsten et al.18 introduced a concept known as energy dissipation circulation function
45
46
47 (EDCF). This function combines the effect of both the maximum energy dissipation in the
48
49 impeller region and the frequency with which the suspended material passes through this region
50
51 (circulation time, tc ). Jüsten et al. examined systematically the influence of stirrer geometry on
52
53
54
the morphology and fragmentation of Penicillium chrysogenum in unaerated vessels and proved
55
56
57
58
59
60 ACS Paragon Plus Environment
5
Industrial & Engineering Chemistry Research Page 6 of 47

1
2
3 d
4 that, at the same specific power input, radial paddle impellers and Rushton turbines with large D
5
6 d
7
ratios caused less hyphae fragmentation in comparison to axial impellers with small ratios and
D
8
9 with low power numbers. In an attempt to find a proper scale-up parameter, it was reported that
10
11 P P 1
12 both the V and the utip approach failed. The energy dissipation circulation function, V ∙ t , which is
I c
13
14
a modified concept of an earlier trial of Smith et al.8, seemed to be the correlation approach with
15
16
17 the best results.18 In the biotechnological field, further studies used the EDCF successfully to
18
19 interpret the effect of agitation on the growth, fragmentation and morphology of
20
21 microorganisms.9,10,31–33 In the field of liquid/liquid dispersions Zhou and Kresta19 were also the
22
23
24 first to consider the drop break-up as the result of the interaction between the local max and the
25
26 circulation time of the mean flow tc .
27
28
29
30
Table 1. Correlations for the estimation of the ratio of maximum to mean energy dissipation rate
31
32 in stirred tanks.
33
34 Research
35 Reference Correlation
method
36
37 Laser
max 1 H D 3
38 Geisler24 = 0.14∙C∙Ne ∙ ( ) ∙ ( )
3 Doppler
39  D d Anemometry
40
max cD D 3 V Liquid/liquid-
41 Liepe34 = π3 ∙ ∙ ( ) ∙ ( 3)
42  Ne d D model system
43 max P ∙V V  Liquid/liquid-
44 McManamey27 = (∙V ) ∙ ( P ) = V with VI = 4 ∙hI ∙d2
 I I system
45 max c
46 = F with
Henzler and  Solid/liquid-
2 2
47 12,14
𝑑 2 2
H -3
48
Biedermann P
.F = P ∙
VD
= (𝐷)
h 3
∙ ( dI) ∙NB 0.6 ∙(sin)1.15 ∙NI 3 ∙ (D) model system
D V
49
50
51
52 Investigations with real systems, such as chemical reactive emulsions or growing
53
54 microorganisms or, remain complex, time- and cost-intensive and affected by multiple (bio-
55
56
57
58
59
60 ACS Paragon Plus Environment
6
Page 7 of 47 Industrial & Engineering Chemistry Research

1
2
3 )chemical parameters. For this reason, the use of model particle systems has been proposed and
4
5
6 reported widely in literature, to facilitate investigations towards a more cost-effective
7
8 optimization of the stirring parameters in the reactors of interest.12 The most frequently reported
9
10 model particle systems include enzymes immobilized on carriers12, oil (or
11
12
13
solvent)/water/surfactant emulsions13,26,28,35–40 and a clay/polymer flocculation system14,29,41–43.
14
15 These systems have been used to investigate mechanical stress phenomena in stirred tanks29,44,
16
17 shake flasks40, model viscometers14, pumps38 and airlift reactors17,45 in a more cost- and time-
18
19
effective way.
20
21
22 The objective of this work is to investigate the influence of various impeller geometries on the
23
24 particle stress in a stirred tank. To quantify the hydromechanical stress, a liquid/liquid model
25
26 system is developed, and the drop breakage induced by different impeller types is monitored in-
27
28
29 line during the stirring process with an endoscope measuring technique. Apart from the
30
31 conventional impeller shapes, which are widely used in technical applications, modifications of
32
33 the impeller geometries are tested in this study. The goal is to identify geometrical modifications,
34
35
36
which could be used to develop “low shear” agitators, in terms of reduced particle breakage. Five
37
38 newly developed impellers are investigated in this work. The design of these impeller
39
40 configurations is mainly based on the cooperation with EvoLogics GmbH. The various stirrers
41
42
are primarily compared in terms of mean specific energy dissipation rate per mass . In an
43
44
45 attempt to interpret the basic hydrodynamic mechanism, which defines the particle stress,
46
47 literature approaches to estimate the maximum energy dissipation and the circulation frequency
48
49 in the impeller swept volume are applied for the correlation of the experimental results. In this
50
51
52 study the reported results are obtained in the frame of a project with biotechnological inititiave.
53
54 They concern a break-up controlled liquid/liquid dispersion, but can also find application in any
55
56
57
58
59
60 ACS Paragon Plus Environment
7
Industrial & Engineering Chemistry Research Page 8 of 47

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

1
2
3 MATERIALS AND METHODS
4
5
6
In the present study, the developed model system consisted of two liquid phases. A Newtonian
7
8
9 fluid silicon oil M100 (Carl Roth) with density D = 965.4 kg/m3 and dynamic viscosity  = 96.2
10
11 mPa∙s at 22.5 °C was chosen as the dispersed phase. Ultrapure water (  0.055 µS/cm,
12
13
14 PURELAB flex 2 system, ELGA Labwater) with density D = 997.6 kg/m3 and dynamic
15
16 viscosity   0.95 mPa∙s at 22.5 °C was used as the continuous phase. The volume phase fraction
17
18 of the dispersed phase was D = 1.5% v/v. To avoid coalescence of the droplets the nonionic
19
20
21 surfactant Triton X-100 (Merck, purity 99.8%) was added at a concentration of 0.232 mmol/L
22
23 (  0.15% w/w with respect to water), which corresponded to a concentration ten times higher
24
25
than the critical micelle concentration of the surfactant/water system (CMC = 0.0232 mmol/L).
26
27
28 The interfacial tension between the phases was measured to be   5.08 mN/m. The dynamic
29
30 viscosities of the single phases were determined using a cone-plate rheometer (MCR 302, Anton
31
32
Paar) with temperature control mechanism. The density measurements were conducted using the
33
34
35 oscillating U-tube principle (DSA 5000 M, Anton Paar). The measurements of the interfacial
36
37 tension between the phases were carried out according to the Pendant drop method (OCA 20,
38
39 Data Physics).
40
41
42 Experiments were performed in a cylindrical glass tank with a dished bottom (Figure 1), an
43
44 H
inner diameter of D = 160 mm and the filling height = 1. Fourteen impellers with various
45 D
46
47 dimensions and geometries were investigated (Table 2, Figure 2). These included common
48
49 impeller types such as Rushton turbines, pitched blade turbines and propellers and five newly
50
51 developed impellers; a propeller with an incorporated circumferential ring and four patent
52
53
54 pending impellers of EvoLogics GmbH (bionic-loop impeller and wave-ribbon impellers). The
55
56 blade length and the disc diameter of the employed Rushton turbines were not designed
57
58
59
60 ACS Paragon Plus Environment
9
Industrial & Engineering Chemistry Research Page 10 of 47

1
2
3 according to the standard dimensions usually reported in literature, but based on the dimensions
4
5
6 of Rushton turbines employed in a prototype lab-fermenter. The off-bottom clearance hc , defined
7
8 as the distance between the centerline of the impeller blades and the bottom of the tank impeller,
9
10 was kept constant at hc = 0.33∙D. The vessel was fitted with four equally spaced stainless-steel
11
12
13
baffles connected at their bottom with a ring. The baffles had a width BB = 12 mm, thickness
14
15 TB = 2 mm, immersed length HB = 114 mm and a baffle ring of BRB = 10 mm width. The distance
16
17 of the baffles from the tank walls was AB ≈ 5 mm. The absence of baffling below the impeller
18
19
plane was a characteristic feature of the employed set-up, with the goal of geometrically
20
21
22 simulating a prototype lab-fermenter. The temperature in the system was controlled at 22.5 °C
23
24 using an external thermostat.
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47 Figure 1. Experimental set-up and dimensions of the stirred tank and basic impeller dimensions.
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
10
Page 11 of 47 Industrial & Engineering Chemistry Research

1
2
3 Table 2. Geometric dimensions and power number (turbulent flow regime) for the investigated
4
5
6 impellers.
7
8 d h hI t Additional
9
Impeller Abbreviated symbol [-] NB [-]  [°] [-] [-] [-] Νe [-]
D d d d dimension
10 RT-d/D=0.6 0.6 6 90 0.2 0.2 0.02 4.2 For RT impellers:
11
Rushton RT-d/D=0.4 0.4 6 90 0.2 0.2 0.02 4.06 b d
= 0.3, dd = 0.54
12 d
13 turbines td
RT-d/D=0.33 0.33 6 90 0.2 0.2 0.02 3.52 = 0.02
14 d
15 PBT-6×90° 0.33 6 90 0.24 0.24 0.02 4.07 -
16 pitched blade PBT-6×45° 0.33 6 45 0.24 0.17 0.02 1.53 -
17 turbines PBT-6×45°-h/d=0.2 0.33 6 45 0.2 0.14 0.02 1.47 -
18
19
PBT-6×22.5° 0.33 6 22.5 0.24 0.09 0.02 0.51 -
20 PROP-h/d=0.28 0.33 3 25 0.28 0.12 0.02 0.3 -
propellers
21 PROP-h/d=0.33 0.325 3 25 0.33 0.14 0.02 0.34 -
22 Ring dimensions:
23 propeller- hR
24 PROPRing-h/d=0.33 0.338 3 25 0.33 0.15 0.02 0.36 = 0.15,
ring impeller d
25 tR
= 0.04
26 d
27 Ring dimensions:
28 bionic-loop hR
= 0.05,
BiLOOP 0.33 7 55 0.24 0.19 0.02 0.7 d
29 impeller tR
30 = 0.02
d
31 WRI-d/D=0.33-51.2° 0.33 - 51.2 - 0.49 0.01 0.52 -
32 wave-ribbon
WRI-d/D=0.4-41.4° 0.4 - 41.4 - 0.35 0.01 0.33 -
33 impellers
34 WRI-d/D=0.4-28.8° 0.4 - 28.8 - 0.25 0.01 0.23 -
35
36
37 An endoscope measurement technique (SOPAT GmbH) was used for the in-line detection of
38
39 the transient and steady state drop size distributions. Two endoscope probes were used in order
40
41 to cover a wide range of drop sizes. The endoscope tip was positioned at the stirrer level,
42
43 hend
44 H
≈ 0.3 in the middle of the two baffles at a distance of approximately 2 cm from the tank wall.
45
46 hend
47 Measurements at different regions of the tank H
≈ 0.8 and 0.65 with respect to the bottom of
48
49 the tank and close to the shaft were also conducted. The deviation of the mean Sauter diameter
50
51
d32 measured at the different tank regions was up to 5%, which is well within the experimental
52
53
54 error, showing spatial homogeneity of the dispersive system in terms of drop sizes inside the
55
56 tank.
57
58
59
60 ACS Paragon Plus Environment
11
Industrial & Engineering Chemistry Research Page 12 of 47

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

1
2
3 For the enhancement of the image quality a mirror was attached to the endoscope tip with a
4
5
6 gap size of 5 mm. The spherical shape of the produced drops allowed a simple determination of
7
8 their size through an automated image analysis software (SOPAT GmbH). For each
9
10 measurement point at least 2500 drops were analyzed. The minimal detectable drop size was
11
12
13 ~10 m. The reproducibility of the experiments was exemplarily tested for ten operating
14
15 conditions, revealing deviations in d32 from 0.4-7.2%.
16
17 The experiments had a duration of 1.5-6 h depending on the time needed to reach a steady state
18
19 1 i
d32 d32,i -5 ∑j=i-5 d32,j m
20
condition, which was defined as ( )
t steady state
= ≤ 0.06 . The minimal stirrer
21 ti -ti-5 min
22
23 speed was chosen to ensure fully developed dispersion (by means of visual observation) and the
24
25
26 maximal stirrer speed was limited to prevent air entrainment from the surface. The absence of
27
28 drop coalescence was confirmed experimentally by lowering the agitation speed drastically after
29
30 a long time of stirring. The drop sizes remained unaffected, even 2 h after stirring at low
31
32
33 rotational frequency, proving that drops do not coalesce in the system within this timescale.
34
35 All experiments were carried out in the turbulent regime, where the power number Ne of an
36
37 impeller is constant. The measurements of the power numbers were conducted with ultrapure
38
39
40
water in the same vessel. For these measurements, the stirrers were attached to a viscosimeter
41
42 (HAAKE Viscotester VT 550, Thermo Fischer Scientific GmbH) to measure torque and
43
44 rotational speed. The agitation power and the power number Ne were calculated as following:
45
46 2∙∙N ∙Μ P
47 Ne = = (3)
∙Ν3 ∙d5 ∙Ν3 ∙d5
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
13
Industrial & Engineering Chemistry Research Page 14 of 47

1
2
3 RESULTS AND DISCUSSION
4
5
6
7 The following sections focus at first on the transient and steady-state drop size distributions of
8
9 the examined liquid/liquid-system. A suitable characteristic drop diameter is chosen as reference
10
11 parameter to quantitively describe the influence of the impeller type on the dispersion process
12
13
14
and the drop breakage. Then, the detailed investigation of the effect of the impeller geometry on
15
16 drop breakage is presented, by using the mean power input per unit mass as a base for the
17
18 comparison among the impellers. Finally, the literature approaches correlating the particle
19
20
breakage with the maximum energy dissipation and the impeller circulation frequency are
21
22
23 applied to the obtained results in order to interpret the main hydrodynamic mechanism for the
24
25 drop breakage.
26
27 Drop Size Distributions
28
29
30 The first goal of this study was to determine the appropriate characteristic drop size, which is
31
32 used to describe breakage kinetics and allow the comparison of the drop deformation caused by
33
34 different impeller types. According to Hinze46, in a break-up controlled dispersive system the
35
36
37
maximum stable drop diameter dmax solely depends on the maximum local energy dissipation
38
39 rate max and can be used to indirectly assess the hydromechanical stress in turbulent systems.40,47
40
41 By accepting that the equilibrium Sauter mean diameter d32 is proportional to dmax , most of the
42
43
stirred tank reactor studies address d32 as the characteristic size of dispersion processes.28,38,48–50
44
45
46 In this study, the evolution of the drop size distributions is analyzed in order to fully comprehend
47
48 the working dispersive system and the break-up mechanisms and to objectively choose an
49
50 appropriate drop diameter as criterion for the comparison of the various impeller types.
51
52
53 Figure 3 shows the development of the characteristic drop diameters over time for
54
55 experiments with the Rushton turbine RT-d/D=0.33. The break-up process is clearly depicted by
56
57
58
59
60 ACS Paragon Plus Environment
14
Page 15 of 47 Industrial & Engineering Chemistry Research

1
2
3 the maximum drop diameter dmax , as well as by the various volume based drop diameter
4
5
6 percentiles (dv10 , dv50 , dv90 ) and the Sauter mean diameter d32 . The number based drop sizes
7
8 (dn10 , dn50 , dn90 ) and the arithmetic mean diameter d10 do not evolve significantly over time.
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
Figure 3. Evolution of drop diameters during the experiment (RT-d/D=0.33, =0.24 W/kg).
32
33
34 To interpret this physical behavior, the cumulative number distributions need to be examined.
35
36 Figure 4(a) and Figure 5(a) depict exemplary number based cumulative distributions at different
37
38 points in time of the experiment and under the variation of . Figure 4(a) shows that since the
39
40
41 earliest experimentation stages more than 60% of the droplets have a diameter of dP ≤ 50 m,
42
43 while up to 90% of the counted droplets have a diameter of dP ≤ 100 m. Over time, as a result
44
45
of the breakage process and the lack of coalescence, the absolute number of the droplets dP ≤
46
47
48 100 m increases significantly, but their percentage changes only slightly. Therefore, the
49
50 changes in the number based percentiles (dn10 , dn50 , dn90 , d10) during the experiment are
51
52
53 insignificant. Figure 5(a) presents the steady state Q0 distributions as a function of  for the
54
55 Rushton turbine RT-d/D=0.33. It is remarkable again that the increase of  makes the Q0 rise less
56
57
58
59
60 ACS Paragon Plus Environment
15
Industrial & Engineering Chemistry Research Page 16 of 47

1
2
3
steeply at first (dP ≤ 50 m) and thus surprisingly the percentage of small droplets is smaller for
4
5
6 higher ̅. By observing the slope of the distribution curves, it can be seen that the steepness of the
7
8 Q0 increases with increasing  for dP > 50 m, but it is only at dP ~100 m that the curves
9
10
11 intersect. These findings indicate a mechanism where the breakage of large drops produces a
12
13 greater number of secondary droplets in the range of 50 m ≤ dP ≤ 150 m, while small droplets
14
15
16
of dP ≤ 50 m are steadily produced and are stable due to the presence of surfactant. Similar
17
18 development of the Q0 functions were reported by Wille et al.28 in a study with a coalescence
19
20
inhibited liquid/liquid system. Moreover, these findings verify the conclusion of Zhou and
21
22
23 Kresta37, who reported that the arithmetic mean diameter d10 was inappropriate as characteristic
24
25 size to relate the droplet size with the turbulent flow created by different impeller types.
26
27
At the same time, the small/medium size droplets (dP ≤ 100 m) dominating the Q0
28
29
30 distributions, are almost negligible for the volume based Q3 distributions. The latter are strongly
31
32
33 affected by the presence of large drops, since the volume is proportional to ~ dP 3 . Over time
34
35 (Figure 4(b)) and by increasing  (Figure 5(b)), the Q3 distributions move to smaller droplets,
36
37 recording the break-up of the big drops present in the dispersion system. These large drops as
38
39
40 shown in Figure 4(a) and Figure 5(a) correspond only to 10% of the total particle number and
41
42 the change of their number could not be depicted clearly in the number distributions. Thus, only
43
44 volume based drop diameter percentiles can be used to describe the influence of the power input
45
46
47 and impeller type on the drop breakage.
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
16
Page 17 of 47 Industrial & Engineering Chemistry Research

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23 Figure 4. Transient cumulative distribution of (a) number and (b) volume for the RT–d/D=0.33
24
25 at =0.24 W/kg.
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47 Figure 5. Cumulative distribution of (a) number and (b) volume at steady state for different
48
49
50 mean specific energy dissipation rates  for RT-d/D=0.33.
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
17
Industrial & Engineering Chemistry Research Page 18 of 47

1
2
3 Self-similarity of drop size distributions
4
5
6 The shape of the drop size distribution provides significant insight into the nature of
7
8 microprocesses involved in forming the distributions. The self-similarity of drop size
9
10 distributions means that distributions follow an invariant pattern and have similar shapes with
11
12
13
respect to dispersion time, agitation speed or any other variable of interest. This invariance can
14
15 be revealed by a normalization of the distributions, often based on d32 and dmax.40,47,50–52 To
16
17 compare the self-similarity of the drop size distributions in the present work various criteria are
18
19 0 0
20
used as displayed in Figure 6. The diagrams show the related standard deviation ratios ,
dn50 d32
21
22 3
and d32
for the steady state drop size distributions of all impellers. For each impeller the whole
23
24
25 operating range of mean specific energy dissipation  is taken into consideration and presented as
26
27 an average single point with the respective deviation bars representing the deviations of self-
28
29 0
30 similarity for different stirring frequencies. In Figure 6(a) the strong deviations of dn50
make
31
32 apparent that the number based distributions are not self-similar. Large deviations are observed
33
34
35 for each impeller separately at the different operating conditions, as well as among the impellers.
36
37 Therefore, no correlation can be established between the number based drop sizes for the various
38
39 impellers. Nevertheless, when the volume based size d32 is used, the ratio of the standard
40
41
42
deviation of the number distributions and the Sauter diameter of the various impellers is constant
43 
44 with an average value of d 0 = 0.36 ± 10%. At the same time, Figure 6(b) clearly proves the self-
32
45
46 3
47 similarity of the volume based distribution with an average value of d32
= 0.35 ± 10%. The
48
49 aforementioned average values derive only from the first eleven impellers in the order presented
50
51
52 in Figure 6. The wave-ribbon impellers are excluded from the calculation of the average values.
53
54 Their distributions show systematically larger deviations from all other impeller types, indicating
55
56
57
58
59
60 ACS Paragon Plus Environment
18
Page 19 of 47 Industrial & Engineering Chemistry Research

1
2
3 that the dispersion mechanism and the microprocesses in the reactor induced by these impellers
4
5
6 probably differ. By examining the widths of the number and volume based distributions, similar
7
8 dn90 -dn10
conclusions were reached. Lack of self-similarity was also observed when the span0: of
9 dn50
10
11 dv90 -dv10
12 the number based distributions was considered, while the span3: dv50
indicated that the volume
13
14 based distributions are equally wide for the various impellers (excluding again the wave-ribbon
15
16
17 stirrers).
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38   
39 Figure 6. Related standard deviation (a) d 0 , d 0 and (b) d 3 of all steady state drop size
n50 32 32
40
41
42
distributions in the whole operating range of mean specific energy dissipation rates  for the
43
44 investigated impellers.
45
46
47 d32
Finally, Figure 7 shows a constant average value of dmax
= 0.54 ± 10% for the different
48
49
50 operating conditions and impellers apart from the wave-ribbon impellers, which is well in
51
52 agreement with literature data in the range of 0.56-0.61 for break-up controlled dispersion
53
54
55
systems.2,47,48 It is concluded that in the investigated coalescence inhibited dispersion system, the
56
57
58
59
60 ACS Paragon Plus Environment
19
Industrial & Engineering Chemistry Research Page 20 of 47

1
2
3 hydromechanical stress caused by various impellers can be quantified only by examining volume
4
5
6 based drop size distributions. The Sauter mean diameter d32 , as a volume based averaged size,
7
8 emphasizes the changes of the large droplets with the time, the power input and the impeller
9
10 type. For all further correlations the Sauter mean diameter d32 is used as reference size.
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31 d
Figure 7. Ratio d 32 of all steady state drop size distributions in the whole operating range of
32 max
33
34 mean specific energy dissipation rates  for the investigated impellers.
35
36
37 Comparative study of impeller type and geometry
38
39
In this section, the steady state Sauter mean diameter d32 is plotted over the mean specific energy
40
41
42 dissipation rate  for the various impeller types. The correlation function between these two
43
44 -b
45
physical sizes has the form: d32 ∝  . Based on the model of Shinnar and Church15, Henzler12
46
-1/3
47 and Wollny36 transformed the correlation as following: d32 ∝  for the particle stress in
48
49 -2/5
50 dissipation range and d32 ∝  for the particle stress in inertial range. The models are valid if
51
52 the flow field is fully turbulent. For fully developed turbulence, the following condition must be
53
54 
55 satisfied:  150...250, with  being the macroscale of turbulence.12,34

56
57
58
59
60 ACS Paragon Plus Environment
20
Page 21 of 47 Industrial & Engineering Chemistry Research

1
2
3 To characterize the predominant forces (inertial, viscous) of the particle stress in the system
4
5
6 under investigation, the determined drop sizes are compared with the size of Kolmogorov
7
8 microeddy  under the various operating conditions in the stirred tank. The calculation of 
9
10 prerequisites the estimation of max for every impeller. To estimate this, the equation of Liepe34
11
12
13
as displayed in Table 1 is used. Values for the factor cD of the formula are extracted from
14
15 literature.34,36 However, data are not available for all impeller types employed here and thus  is
16
17 estimated only for some of the impellers. The calculated turbulence parameters are given in
18
19 d32
20 Table S1. In the majority of cases the condition of

< 6 is satisfied, indicating particle stress
21
22
in the dissipation range of microturbulence and dominant viscous stresses. For this reason, the
23
24
-1/3
25 experimental data in the following diagrams are plotted using the correlation d32 ∝  to which
26
27
they fit at a remarkable degree.
28
29
d
30 In Figure 8(a) the effect of the impeller diameter ratio on the droplet size is presented. The
D
31
32
33 Sauter mean diameter d32 is plotted versus the mean specific energy dissipation rate  for three
34
35 d
Rushton turbines with varying diameter. At equal  the impellers with larger D ratios produce less
36
37
38 breakage of the dispersed droplets, which confirms the results reported in relevant research
39
40 works.12,14,18,36 At equal  a larger impeller operates at a significantly lower rotational frequency
41
42
43
and thus at a lower tip speed, which apparently plays an important role in the drop break-up.
44
45 Zhou and Kresta53 conducted LDA measurements, showing that the impeller diameter has a great
46
47 effect on the maximum turbulent energy dissipation rate max . The authors reported a change of
48
49 max
50 the turbulent flow field with increasing impeller diameter and a decrease of , attributed to

51
52 max
strong interactions between the impeller and the tank walls.53 The lower values for larger
53 
54
55 impellers are assumed to be the reason for the reduced particle stress.
56
57
58
59
60 ACS Paragon Plus Environment
21
Industrial & Engineering Chemistry Research Page 22 of 47

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23 d
Figure 8. Influence of (a) the impeller diameter ratio D and (b) the stirrer blade angle on the
24
25
26 steady state drop size d32 in correlation with the mean energy dissipation rate .
27
28
29 Figure 8(b) shows the effect of the impeller blade angle  on the drop size for pitched blade
30
31
32
turbines. The only varying parameter for the stirrers was the blade angle, while the ratios of the
33
d h t
34 other dimensions D, d, were kept constant. The decrease of the blade angle corresponds to a
35 d
36
37 significant decrease in power number (Table 2), which further results in increased droplet
38
39 breakage when equal power input per unit mass is applied. The results agree with the work of
40
41 Henzler and Biedermann12,14, who considered that the increase in the angle of the impeller blades
42
43
44 reduces the particle stress. The PBT-6×22.5° and PBT-6×45° are axial flow impellers, while the
45
46 h
impeller PBT-6×90°, which is in fact a flat paddle impeller with small ratio, causes an
47 d
48
49 important change in the flow direction, which is expected to become radial in this case.
50
51 Apparently, the radial induced flow is beneficial for reduced particle stress. Ranade and Joshi54
52
53
54
investigated the effect of impeller blade pitch (=30°, 45° and 60°) on the flow pattern by means
55
56 of LDA. They established that the blade angle significantly affects the flow characteristics, since
57
58
59
60 ACS Paragon Plus Environment
22
Page 23 of 47 Industrial & Engineering Chemistry Research

1
2
3 an increased  (up to 60°) results in an enlarged average liquid velocity, increased total flow
4
5
6 circulation, and higher turbulence intensity. However, these findings cannot be directly
7
8 correlated with the results of the present study to which they seem contradictory.
9
10 In Figure 9 various propeller types are compared in terms of drop break-up. A comparison of
11
12
13
the two conventional 3-blade propellers (PROP-h/d=0.28 and PROP-h/d=0.33) with different
14
h
15 blade height shows that an increase in the impeller height leads to an increase in the power
16 d
17
18 number of the agitator and thus to reduced drop breakage at equal .
19
20 Newly developed propeller types were also investigated in this study. The bionic loop impeller
21
22
BiLOOP, as well as the wave-ribbon configurations WRI-d/D=0.33-51.2°, WRI-d/D=0.4-41.4°,
23
24
25 WRI-d/D=0.4-28.8° are compared with the conventional 3-blade propellers. The BiLOOP, which
26
27 is here operated in an up-pumping mode, forms relatively larger drops than the PROP-h/d=0.33
28
29 and PROP-h/d=0.28 at equal . The wave-ribbon impellers produce the greatest drop breakage
30
31
32 and thus the largest interfacial area per volume. Therefore, they appear to be particularly suited
33
34 for dispersion processes, where a large droplet surface is required. Nevertheless, they could not
35
36 accommodate the requirements of a “low shear” demanding process, where the particle breakage
37
38
39 is undesired. Additionally, the wave ribbon impeller WRI-d/D=0.33-51.2° causes less drop
40
d
41 break-up than the wave ribbon impellers with larger diameter ratios D, leading to the conclusion
42
43
44 that the angle of the band is more crucial than the impeller diameter.
45
46 Finally, the conventional 3-blade propeller PROP-h/d=0.33 was modified by adding a ring to
47
48 connect the blades. The addition of the circumferential ring has a small impact on the power
49
50
51 number of the impeller (Ne=0.34 for PROP-h/d=0.33 and Ne=0.36 for PROPRing-h/d=0.33
52
53 respectively). Nevertheless, stirring with the propeller PROPRing-h/d=0.33 results in droplet size
54
55 distributions with significantly larger droplet diameters. This propeller type seems to cause the
56
57
58
59
60 ACS Paragon Plus Environment
23
Industrial & Engineering Chemistry Research Page 24 of 47

1
2
3 lowest particle stress of all examined propeller shapes. It is assumed that due to the addition of
4
5
6 the ring the supplied energy dissipates in a more equally distributed way in the impeller region,
7
8 decreasing the shear strength of the agitator. These results show that a modification of
9
10 conventional propellers with the usage of circumferential rings could probably change the flow
11
12
13
field in a significant way and consequently the hydromechanical stress acting on particulate
14
15 material. It is estimated that modifications in this direction could lead to an impeller geometry
16
17 suited for applications where “low shear” stirring conditions are desired.
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38 Figure 9. Influence of the propeller type on the steady state drop size d32 in correlation with the
39
40
41 mean energy dissipation rate .
42
43
44 The comparison of the various impeller geometries investigated in the present work is depicted
45
46 in Figure 10. The radial impeller types, i.e., Rushton turbines and the PBT–6×90° (which is
47
48 h
49 practically a type of flat paddle impeller with small ratio), cause less drop breakage at equal
d
50
51 specific power input in comparison to the axial propellers (pitched blade turbines, propellers) and
52
53
54 the modified bionic-loop and wave-ribbon impellers. Simultaneously, the drop breakage caused
55
56
57
58
59
60 ACS Paragon Plus Environment
24
Page 25 of 47 Industrial & Engineering Chemistry Research

1
2
3 by the axial 6×45°-pitched blade impellers and the modified propeller with the peripheral ring
4
5
6 PROPRing-h/d=0.33 is low and comparable with the one produced by the radial impellers.
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28 Figure 10. Influence of the impeller type on the steady state drop size d32 in correlation with the
29
30 mean energy dissipation rate .
31
32
33
34
The power input range of operation for the different impellers varies significantly. The radial
35
36 impeller types can ensure a complete oil dispersion at lower power inputs. An increase in the
37
38 impeller diameter d for the Rushton turbines, as well as in the blade angle  for the pitched blade
39
40
41
turbines result in a fully developed droplet dispersion at lower . All four modified stirrers
42
43 (PROPRing-h/d=0.33 and WRIs) are especially characterized by a limited operational area of ,
44
45 providing a full dispersion only at high power inputs and causing surface aeration shortly after
46
47
that point.
48
49
50 In the past, the axial flow impellers had been considered as “low shear” agitators due to their
51
52 low power numbers.25 However, the studies in particulate systems reveal that the particle stress
53
54 caused by these impellers is higher.13,14,26,28,36 This statement is in general agreement with the
55
56
57
58
59
60 ACS Paragon Plus Environment
25
Industrial & Engineering Chemistry Research Page 26 of 47

1
2
3 results of the present study with the exceptions of PBT-6×45° and PROPRing-h/d=0.33. The
4
5
6 interpretation of the increased hydromechanical stress caused by axial impellers was proven to
7
8 be challenging. Due to their low power numbers, the axial stirrers need to rotate at higher speed
9
10 in order to operate at equal specific power input with the radial ones. Nevertheless, at the same
11
12 max
13 time fluid dynamic investigations correlate the axial impellers with lower ratios. Which fluid

14
15 dynamic effect is responsible for the high hydromechanical stress of the conventional axial
16
17
18 impellers is not yet clear. Wollny36 tried to enlighten this research topic by means of
19
20 Computational Fluid Dynamics (CFD). He compared the flow profiles produced by a radial
21
22 Rushton turbine and an axial 3×24°-pitched blade turbine. According to the conducted
23
24
25
simulations, the significant region for the particle stress corresponds to a volume smaller than
26
27 1% of the total reactor volume and is limited in the impeller zone and. The volume of the
28
29 impeller zone as estimated by Wollny was in fact of similar magnitude with the impeller swept
30
31
volume as defined by McManamey27 (Table 1). This volume, which is affected by high velocity
32
33
34 gradients, was found to be smaller for the axial impeller. At equal mean specific power input, the
35
36 shear gradients in the impeller zone were significantly bigger for the axial pitched blade agitator.
37
38 The amount of elongation gradients barely differed for the two impeller types. It was concluded
39
40
41 that the axial flow impellers are characterized by a smaller volume where the particle
42
43 deformation can occur, but also by higher shear gradients in this region, which leads to an
44
45 increased particle stress in comparison to the radial impellers. Wille et al. 28 also concluded that
46
47
48
axial flow impellers with low power numbers cause higher particle stress than radial stirrers.
49
50 However, Wille et al.13 and Langer et al.42 considered as the predominant cause of the particle
51
52 deformation and breakage the elongation gradients attributed to the axial induced agitators, and
53
54
not the shear gradients as Wollny described. Langer et al.42 showed experimentally that in the
55
56
57
58
59
60 ACS Paragon Plus Environment
26
Page 27 of 47 Industrial & Engineering Chemistry Research

1
2
3 macroscopic flow field the elongation flow can cause stronger particle breakage than the shear
4
5
6 flow. Wille et al.13 carried out PIV measurements for a radial Rushton turbine and an axial
7
8 3×24°-pitched blade turbine. At equal specific power input, the axial flow impeller produced a
9
10 considerably high elongation flow. At the same time, for the axial impeller, elongation and shear
11
12
13
gradients extended over the whole volume between the impeller and the bottom of the tank. In
14
15 the case of the radial stirrer, the respective gradients were present only to the vicinity of the
16
17 impeller. It was concluded that the axial impellers build a larger area with high velocity
18
19
gradients. This results in longer residence times of the particles in the corresponding region,
20
21
22 increasing the probability of particle breakage.
23
24 Despite the contradictory approaches concerning the responsible forces for the
25
26 hydromechanical stress (shear, normal stresses) in stirred tanks, the published studies agree on
27
28
29 the fact that low power number axial impellers cause greater break-up on suspended particles.
30
31 However, in the present study, the two 6×45°-pitched blade impellers produce drop size
32
33 distributions very similar to the radial Rushton turbine RT-d/D=0.33 despite the important
34
35
36 difference in the power number among these stirrer types (Table 2). Similar results had been
37
38 reported by Wille et al.28, where a Rushton turbine and a 6×45°-pitched blade turbine produced
39
40 similar equilibrium drop sizes, despite the higher circumferential velocity reached by the pitched
41
42
43
blade impeller in order to operate at equal specific power input.
44
45 On the other hand, the modified propeller PROPRing-h/d=0.33 with the peripheral ring
46
47 appears to form significantly large droplets despite the very low power number. It is estimated
48
49
that the addition of the ring allows a more homogeneous distribution of the energy dissipation in
50
51
52 the impeller region, which reduces the drop breakage. Preliminary investigations of the flow
53
54 field with PIV indicated a significant change in the flow direction by the addition of the
55
56
57
58
59
60 ACS Paragon Plus Environment
27
Industrial & Engineering Chemistry Research Page 28 of 47

1
2
3 peripheral ring. Although the conventional propellers induce an axial flow, the flow field formed
4
5
6 by PROPRing-h/d=0.33 shows radial components as well.
7
8 With respect to the other newly developed impellers, the abovementioned preliminary PIV
9
10 investigations indicated an axial induced flow for WRI-d/D=0.33-51.2° and BiLOOP, an axial
11
12
13
flow with radial components for WRI-d/D=0.4-41.4° and a radial flow with axial components for
14
15 WRI-d/D=0.4-28.8°. The drop breakage caused by BiLOOP and WRI-d/D=0.33-51.2° is
16
17 moderate and equivalent to conventional axial stirrers. On the other hand, the decrease in the
18
19
angle of the wave-ribbon configurations triggers strong breakage leading to very small drop
20
21
22 sizes. It is worth mentioning that the drop size distributions recorded for the wave-ribbon
23
24 impellers do not fulfill the criterion of self-similarity described in Figure 6 and Figure 7. This is
25
26 probably caused by different fluid dynamics, long residence time of drops in regions with high
27
28
29 shear/elongation stresses, or even by impact stress between the drops and the impeller.
30
31 Further comparison with literature-Correlation approaches
32
33 The results in the previous paragraphs confirmed the published effects of agitation on
34
35
36
hydromechanical stress in stirred tanks but also expand them by considering modified and newly
37
38 developed impeller types. It is apparent that in stirred tanks by considering the average specific
39
40 power input the drop breakage depends strongly on the impeller geometry.
41
42
Apart from the mean mass or volume specific power input, other parameters have often been
43
44
45 used in the literature to correlate the experimental results. Most frequently the efforts focus on
46
47 using power input based parameters that can indirectly correlate the particle size with the
48
49 maximum energy dissipation max in the stirred reactor. The most significant of these correlations
50
51
52 are applied here in order to put the present work into the general framework of the currently
53
54 available literature that concerns the selection of impeller type in stirring processes. The Sauter
55
56
57
58
59
60 ACS Paragon Plus Environment
28
Page 29 of 47 Industrial & Engineering Chemistry Research

1
2
3 mean diameters at the steady state are used for all further correlations presented in the following
4
5
6 paragraphs.
7
8 In the attempt to comprehend the effect of max on drop breakage, one must take into
9
10 consideration that the drops are subject to stress mainly in the impeller swept volume VI where
11
12
13 the maximum energy dissipation max occurs. McManamey27 defined the impeller swept volume
14

15 by VI = 4 ∙hI ∙d2 , estimating the maximum energy dissipated in the swept volume of the impeller
16
17 max V 4 V
18 as 
= = ∙
VI  d2 ∙hI
(Table 1). The calculated values of VI for the impellers under investigation
19
20
21 are presented in Table S2. Figure 11 presents the modified data based on VI . It becomes clear
22
P
23 that approaching max by ∙VI
can achieve a general clustering of the experimental data for the
24
25
26 conventional impeller types. With respect to the modified impeller types, the BiLOOP impeller
27
28 correlates well with the other data, while the PROPRing-h/d=0.33 deviates up to 30%. The data
29
30
of the modified wave-ribbon impellers are not in agreemement with the rest of the experimental
31
32
33 results. The estimation of the critical impeller volume as proposed by McManamey has
34
35 limitations when it comes to non-conventional stirrer types. Limitations of the approach have
36
37 been mentioned in literature by other authors as well.14,19, 26
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
29
Industrial & Engineering Chemistry Research Page 30 of 47

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23 Figure 11. Influence of the impeller type on the steady state drop size d32 in correlation with the
24
25 P
26 mean energy dissipation rate per impeller swept volume ∙V .
I
27
28
29
Henzler and Biedermann14 suggested a modification of the calculation of the impeller volume
30
31
32 so that other geometrical parameters of the impeller, e.g., number of blades and blade angle, are
33
34 also taken into account. Therefore, they introduced a geometrical factor F (Table 1) to
35
36 max c
approximate the maximum energy dissipation rate as = F. For the impellers employed in this
37 
38
39 study, the F factors were calculated (Table S2), with the exception of the wave-ribbon impellers
40
41 due to the lack of actual blades in these configurations. In Figure 12 the steady state Sauter mean
42
43 1
44 diameter is plotted as a function of F ∙
̅ . This function correlates the experimental data of the
45
46
conventional impeller types to a good degree but again it fails to predict the results of BiLOOP
47
48
49 and PROPRing-h/d=0.33, which appear to deviate up to 30% and 50% respectively. These
50
51 deviations can be justified by the fact that the F function does not include any term that can
52
53 describe the influence of the circumferential ring. In general, the F function approaches in a
54
55
56 satisfactory way the critical volume around the impeller, where the drop breakage takes place,
57
58
59
60 ACS Paragon Plus Environment
30
Page 31 of 47 Industrial & Engineering Chemistry Research

1
2
3 for conventional stirrer types, but it must be expanded to facilitate the geometrical description of
4
5
6 new, more complex impeller types.
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
1
28 Figure 12. Correlation of the steady state drop size d32 with the F ∙ ̅ for all impellers.
29
30
31
32 Jüsten et al.18 proposed the use of the energy dissipation circulation function (EDCF), defined
33
P 1
34 as V ∙ t , to correlate and predict the breakage of the various impellers. This function considers not
35 I c

36
37 only the maximum energy dissipation in the impeller zone but also the frequency at which the
38
39 particles are present in this zone. Even though the EDCF approach originates and finds
40
41 application mainly in biotechnological systems, similar approaches have been followed also for
42
43
44 liquid/liquid dispersions. Zhou and Kresta19 stated the need to consider simulateneously the local
45
46 1
max and the circulation frequency tc
in order to correlate the data for their employed silicon
47
48
49 oil/water dispersion. In this work, the EDCF approach as defined by Jüsten et al.18 is applied. To
50
51 1
calculate the circulation frequency the approach of Smith et al.8 and Jüsten at al.18 was used,
52 tc
53
54 1 Fl∙N∙d3
55 i.e., tc
= V
, where Fl is the impeller flow number. To accommodate the range of impellers
56
57
58
59
60 ACS Paragon Plus Environment
31
Industrial & Engineering Chemistry Research Page 32 of 47

1
2
3 employed, the flow numbers Fl for the standard impellers were calculated based on the power
4
5
6 numbers, by using formulas available in literature as they have been reviewed by Jüsten et al.18.
7
8 The calculated flow numbers for every impeller are displayed in Table S2. No prediction could
9
10 be made for the modified impeller types.
11
12 P 1
13 Figure 13 presents the clustering of the experimental data as a function of ∙
VI tc
for the
14
15
conventional impellers. By this approach, the correlation of the measured drop sizes for the
16
17
18 various impellers is again satisfactory. The deviations among the stirrers are considered
19
20 insignificant (constantly less than 20%), taking into account that the flow numbers derive from
21
22 general formulas and not from experimental measurements in the specific stirred system. The
23
24
25 good correlation achieved by using the EDCF indicates the importance of the frequency at which
26
27 the drops are subjected to high stresses near the impeller and the need to consider the circulation
28
29 time as key-element for the interpretation of the particle stress in stirred systems.
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52 Figure 13. Steady state drop size d32 as a function of energy dissipation rate per impeller swept
53
54 P 1
volume and circulation frequency V ∙ t .
55 I c

56
57
58
59
60 ACS Paragon Plus Environment
32
Page 33 of 47 Industrial & Engineering Chemistry Research

1
2
3 CONCLUSIONS
4
5
6 This work investigated the effect of the impeller geometry on particle stress in stirred tanks. The
7
8 most widely used conventional agitators and five new impellers were examined in terms of
9
10 particle stress. To quantify the hydromechanical stress in stirred tanks, a non-coalescing
11
12
13
liquid/liquid dispersion system was used.
14
15 The drop breakage was recorded in-line by means of an endoscope measuring technique. To
16
17 describe the influence of the power input and impeller type in breakage only the volume based
18
19
distributions could be used as reference parameters. Therefore, the Sauter mean diameter d32 was
20
21
22 used as reference size under a broad range of operating conditions.
23
24 The comparison of the impellers was based on the mean specific energy dissipation rate ̅ . The
25
26 diameter d of the impeller, the impeller blade angle  and the impeller blade height h were
27
28
29 proven to be important geometrical features of the agitator. An increase of the aforementioned
30
31 geometrical dimensions resulted in reduced particle stress and thus larger drop sizes. The
32
33 impeller type was also decisive for the drop breakage and the particle stress in stirred tanks. The
34
35
36 high power-number, radial impellers caused reduced drop breakage compared to the low power-
37
38 number, axial agitators. Exceptions were the results of the 6×45°-pitched blade turbines and the
39
40 modified propeller type PROPRing-h/d=0.33. These axial impeller types, despite their low
41
42
43
power number, resulted in drop sizes comparable to the ones formed by the radial stirrers. The
44
45 addition of a peripheral ring in the modified propeller must be further examined as a possible
46
47 alternative for the development of a “low shear” impeller, for application where the particle
48
49
breakage is undesired. The modified impellers, designed by EvoLogics GmbH, followed the
50
51
52 general trend of the conventional propellers employed in this work. Especially, the wave-ribbon
53
54 configurations resulted in the smallest drop sizes and could be of high interest in multiphase
55
56
57
58
59
60 ACS Paragon Plus Environment
33
Industrial & Engineering Chemistry Research Page 34 of 47

1
2
3 dispersion applications where the formation of a large interfacial area is desired. In these
4
5
6 configurations, the angle of the band seemed to be a more crucial dimension than the impeller
7
8 diameter and must be considered for further design optimization. All modified impellers were
9
10 additionally characterized by a narrow operational area of power input being limited by the
11
12
13
conditions of fully developed dispersion and surface aeration.
14
15 To interpret the influence of the impeller type on the particle stress, multiple fluid dynamic
16
17 effects need to be considered. The indirect correlation of the drop sizes with the maximum local
18
19
energy dissipation rate max was made. For this purpose, the approaches of McManamey27 and
20
21
22 Henzler and Biedermann14 to estimate the critical impeller swept volume (VI and VD ), where the
23
24 maximum energy dissipation occurs, were followed. Both approaches clustered well the results
25
26 for the conventional impeller types, but they do not match the data for the modified impeller
27
28
29 geometries. The concept of EDCF was applied here, showing that the frequency by which the
30
31 drops are passing through the impeller region must be quantified and used for further
32
33 clarification of the particle stress induced by the impeller in a stirred system. Nevertheless,
34
35
36 further systematic fluid dynamic investigations are needed to fully characterize the size of the
37
38 area where the high velocity gradients occur and the circulation time of the mean flow for the
39
40 conventional as well as the new stirrer types.
41
42
Indisputably, as next research step, the exact flow fields formed by each stirrer, as well as the
43
44
45 mechanisms (shear, elongation forces) responsible for the particle stress, need to be described by
46
47 means of further experimental (PIV) and numerical investigations (CFD) in order to allow a
48
49 more complete interpretation of the results obtained in the present study. Further fluid dynamic
50
51
52 analysis of the new impeller types should be carried out. Finally, the reported results are valid for
53
54 the specific employed set-up, which is characterized by absence of baffling below the impeller
55
56
57
58
59
60 ACS Paragon Plus Environment
34
Page 35 of 47 Industrial & Engineering Chemistry Research

1
2
3 plane. Whether and how the baffle position could affect the particle stress induced by the
4
5
6 examined stirrers should also be a topic of further research.
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
35
Industrial & Engineering Chemistry Research Page 36 of 47

1
2
3 NOMENCLATURE
4
5
6
Symbols used
7
8 AB [mm] Distance of baffles from the tank wall
9 BB [mm] Width of baffles
10 BRB [mm] Width of baffle ring
11 b [mm] Blade length for Rushton turbines
12 C, c [-] Constant factors
13 
14 cD [-] Impeller factor for estimation of max (Liepe34)

15 D [mm] Stirred tank diameter
16 d [mm] Impeller diameter
17
dd [mm] Disk diameter for Rushton turbines
18
19 dp m] Particle size
20 d10 m] Arithmetic mean diameter
21 d32 m] Sauter mean diameter
22 dn10 , dn50 , dn90 m] Number based percentiles
23
24 dv10 , dv50 , dv90 m] Volume based percentiles
25 dmin , dmax m] Minimum and maximum diameter
26 F [-] Geometrical factor (Henzler and Biedermann12,14)
27 Fl [-] Flow number of impeller
28 [mm] Filling height of stirred tank
H
29
30 HB [mm] Immersed length of baffles
31 h [mm] Height of impeller blades
32 hc [mm] Off-bottom clearance
33 hend [mm] Position level of endoscope tip
34 hI (= h ∙ sin ) [mm] Vertical impeller height
35
hR [mm] Height of impeller ring
36
37 M [N∙m] Torque
38 N [rpm], [rps] Rotational frequency
39 NB [-] Number of impeller blades
40 Ne [-] Impeller power (Newton) number
41 NI [-] Number of impellers
42
P [W] Power input
43
44 PD [W] Power input in a defined volume VD
45 Q0 , Q3 [-] Cumulative drop size distribution of number and volume
46 TB [mm] Thickness of baffles
47 t [mm] Thickness of impeller blades
48
t [min], [h] Time
49
50 tc [s] Impeller circulation time
51 td [mm] Disk thickness for Rushton turbines
52 tR [mm] Thickness of impeller ring
53 utip [m/s] Impeller tip speed
54
55
u' [m/s] Velocity of turbulent fluctuation
56 V [m3] Volume of fluid in stirred tank
57
58
59
60 ACS Paragon Plus Environment
36
Page 37 of 47 Industrial & Engineering Chemistry Research

1
2
3 VD [m3] Defined volume around the impeller
4
5
VI [m3], [cm3] Impeller swept volume
6
7 Greek symbols
8  [°] Angle of impeller blades
9  [% w/w] Mass fraction of surfactant with respect to water mass
10
r [m] Distance between two neighboring points in the flow field
11
12 ̅ [W/kg] Mean specific energy dissipation rate
13 loc , max [W/kg] Local and maximum energy dissipation rate
14  [S/cm] Electrical conductivity
15 h
16  (≈ ) [mm] Macroscale of turbulence
2
17
 [mm], [m] Kolmogorov microscale
18
19  [mPa∙s] Dynamic viscosity
20 fl [m2/s] Kinematic viscosity of fluid
21 , D [kg/m3] Density of fluid and disperse phase
22  [mN/m] Interfacial tension
23
Standard deviation of drop size distribution of number and
24 0 , 3 m]
25 volume
26 t [N/m2] Turbulent stress
27 Volume fraction of dispersed phase with respect to total
D [% v/v]
28 fluid volume
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
37
Industrial & Engineering Chemistry Research Page 38 of 47

1
2
3 ASSOCIATED CONTENT
4
5
6 Supporting Information. (Table S1) Estimated ranges of turbulence parameters; (Table S2)
7
8 Impeller swept volume VI, geometrical factor F and flow number Fl for the investigated
9
10 impellers
11
12
13
14
AUTHOR INFORMATION
15
16 Corresponding Author
17
18
19 *Chrysoula Bliatsiou, Chair of Chemical and Process Engineering, Technische Universität
20
21 Berlin, Tel: +49 30 314 25538, E-mail: c.bliatsiou@tu-berlin.de
22
23
24
25
ORCID
26
27 Chrysoula Bliatsiou: 0000-0001-9394-0031
28
29
30 Author Contributions
31
32 The manuscript was written through contributions of all authors. All authors have given approval
33
34
35 to the final version of the manuscript.
36
37
38 Funding Sources
39
40
41
The research was financially supported by the Deutsche Forschungsgemeinschaft (DFG
42
43 SPP1934).
44
45
46 Notes
47
48
49 The authors declare no competing financial interest.
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
38
Page 39 of 47 Industrial & Engineering Chemistry Research

1
2
3 ACKNOWLEDGMENT
4
5
6 This work is part of the Priority Program “Dispersity-, structural, and phase- changes of proteins
7
8 and biological agglomerated in biotechnological processes”. Financial support by the Deutsche
9
10 Forschungsgemeinschaft (DFG SPP1934) is gratefully acknowledged. The authors wish to thank
11
12
13
EvoLogics GmbH for the cooperation and Ms. Lena Hohl for constructive criticism of the
14
15 manuscript.
16
17
18 ABBREVIATIONS
19
20
CFD, Computational Fluid Dynamics; LDA, Laser Doppler Anemometry; PIV, Particle Image
21
22
23 Velocimetry.
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
39
Industrial & Engineering Chemistry Research Page 40 of 47

1
2
3 REFERENCES
4
5
6 (1) Kumar, S.; Ganvir, V.; Satyanand, C.; Kumar, R.; Gandhi, K. S. Alternative mechanisms
7
8 of drop breakup in stirred vessels. Chem. Eng. Sci. 1998, 53 (18), 3269–3280.
9
10
11 (2) Giapos, A.; Pachatouridis, C.; Stamatoudis, M. EFFECT OF NUMBER OF IMPELLER
12
13
14
BLADES ON THE DROP SIZES IN AGITATED DISPERSIONS. Chem. Eng. Res. Des.
15
16 2005, 83 (A12), 1425–1430.
17
18
19 (3) Maaß, S.; Metz, F.; Rehm, T.; Kraume, M. Prediction of drop sizes for liquid–liquid
20
21
systems in stirred slim reactors—Part I: Single stage impellers. Chem. Eng. J. 2010, 162
22
23
24 (2), 792–801.
25
26
27 (4) Boxall, J. A.; Koh, C. A.; Sloan, E. D.; Sum, A. K.; Wu, D. T. Droplet Size Scaling of
28
29 Water-in-Oil Emulsions under Turbulent Flow. Langmuir 2012, 28, 104–110.
30
31
32
33 (5) Roudsari, S. F.; Dhib, R.; Ein-Mozaffari, F. Impact of Impeller Type on Methyl
34
35 Methacrylate Emulsion Polymerization in a Batch Reactor. J. Appl. Polym. Sci. 2014, 131,
36
37 40496.
38
39
40
41
(6) Martín, M.; Montes, F. J.; Galán, M. A. Influence of Impeller Type on the Bubble
42
43 Breakup Process in Stirred Tanks. Ind. Eng. Chem. Res. 2008, 47 (16), 6251–6263.
44
45
46 (7) Amanullah, A.; Buckland, B. C.; Nienow, A. W. Mixing in the Fermentation and Cell
47
48
Culture Industries. In Handbook of Industrial Mixing; Paul, E. L., Atiemo-Obeng, V. A.,
49
50
51 Kresta, S. M., Eds.; John Wiley & Sons, Inc.: Hoboken, New Jersey, 2003; pp 1071–1170.
52
53
54 (8) Smith, J. J.; Lilly, M. D.; Fox, R. I. The Effect of Agitation on the Morphology and
55
56
57
58
59
60 ACS Paragon Plus Environment
40
Page 41 of 47 Industrial & Engineering Chemistry Research

1
2
3 Penicillin Production of Penicillium Chrysogenum. Biotechnol. Bioeng. 1990, 35, 1011–
4
5
6 1023.
7
8
9 (9) Jüsten, P.; Paul, G. C.; Nienow, A. W.; Thomas, C. R. Dependence of Penicillium
10
11 Chrysogenum Growth, Morphology, Vacuolation, and Productivity in Fed-Batch
12
13
14
Fermentations on Impeller Type and Agitation Intensity. Biotechnol. Bioeng. 1998, 59 (6),
15
16 762–775.
17
18
19 (10) Amanullah, A.; Jüsten, P.; Davies, A.; Paul, G. C.; Nienow, A. W.; Thomas, C. R.
20
21
Agitation induced mycelial fragmentation of Aspergillus Oryzae and Penicillium
22
23
24 Chrysogenum. Biochem. Eng. J. 2000, 5 (2), 109–114.
25
26
27 (11) Liu, J.; Svärd, M.; Rasmuson, Å. C. Influence of Agitation on Primary Nucleation in
28
29 Stirred Tank Crystallizers. Cryst. Growth Des. 2015, 15 (9), 4177–4184.
30
31
32
33 (12) Henzler, H.-J. Particle Stress in Bioreactors. In Influence of Stress on Cell Growth and
34
35 Product Formation; Schügerl K. et al., Eds.; Advances in Biochemical
36
37 Engineering/Biotechnology, Vol. 67; Springer, Berlin, Heidelberg, 2000; pp 35-82
38
39
40
41
(13) Wille, M.; Langer, G.; Werner, U. The Influence of Macroscopic Elongational Flow on
42
43 Dispersion Processes in Agitated Tanks. Chem. Eng. Technol. 2001, 24 (2), 119–127.
44
45
46 (14) Henzler, H.-J.; Biedermann, A. Modelluntersuchungen zur Partikelbeanspruchung in
47
48
Reaktoren (Model Studies on Particle Stress in Reactors). Chem. Ing. Tech. 1996, 68 (12),
49
50
51 1546–1561.
52
53
54 (15) Shinnar, R.; Church, J. M. Statistical Theories of Turbulence in Predicting Particle Size in
55
56
57
58
59
60 ACS Paragon Plus Environment
41
Industrial & Engineering Chemistry Research Page 42 of 47

1
2
3 Agitated Dispersions. Ind. Eng. Chem. 1960, 52 (3), 253–256.
4
5
6
7 (16) Kolmogorov, A. N. The Local Structure of Turbulence in Incompressible Viscous Fluid
8
9 for Very Large Reynolds Numbers. Proc. Math. Phys. Sci. 1991, 434 (1890), 9–13.
10
11
12 (17) Mahnke, E. U.; Büscher, K.; Hempel, D. C. A Novel Approach for the Determination of
13
14
15
Mechanical Stresses in Gas-Liquid Reactors. Chem. Eng. Technol. 2000, 23 (6), 509–513.
16
17
18 (18) Jüsten, P.; Paul, G. C.; Nienow, A. W.; Thomas, C. R. Dependence of Mycelial
19
20 Morphology on Impeller Type and Agitation Intensity. Biotechnol. Bioeng. 1996, 52 (6),
21
22
672–684.
23
24
25
26 (19) Zhou, G.; Kresta, S. M. Correlation of mean drop size and minimum drop size with the
27
28 turbulence energy dissipation and the flow in an agitated tank. Chem. Eng. Sci. 1998, 53
29
30 (11), 2063–2079.
31
32
33
34 (20) Kresta, S. Turbulence in Stirred Tanks: Anisotropic, Approximate, and Applied. Can. J.
35
36 Chem. Eng. 1998, 76 (3), 563–576.
37
38
39 (21) Sheng, J.; Meng, H.; Fox, R. O. A large eddy PIV method for turbulence dissipation rate
40
41
42
estimation. Chem. Eng. Sci. 2000, 55 (20), 4423–4434.
43
44
45 (22) Wu, H.; Patterson, G. K. Laser-Doppler Measurements of Turbulent-Flow Parameters in a
46
47 Stirred Mixer. Chem. Eng. Sci. 1989, 44 (10), 2207–2221.
48
49
50
(23) Cutter, L. A. Flow and Turbulence in a Stirred Tank. AIChE J. 1966, 12 (1), 35–45.
51
52
53
54 (24) Geisler, R. Turbulente Schubspannung und Hydrodynamische Partikelbeanspruchung im
55
56 Rührkessel. In 10. Köthener Rührerkolloqium; Proceedings of 10. Köthener
57
58
59
60 ACS Paragon Plus Environment
42
Page 43 of 47 Industrial & Engineering Chemistry Research

1
2
3 Rührerkolloqium, Köthen, 2007; Sperling, R., Ed.; Hochschule Anhalt: Köthen, 2007.
4
5
6
7 (25) Nienow, A. W. The Many Sheer or (Shear?) Myths of Bioprocessing. In 14. Kötherner
8
9 Rührerkolloqium; Proceedings of 14. Köthener Rührerkolloqium, Köthen, 2011; Sperling,
10
11 R., Ed.; Hochschule Anhalt: Köthen, 2011.
12
13
14
15
(26) Pacek, A. W.; Chamsart, S.; Nienow, A. W.; Bakker, A. The influence of impeller type on
16
17 mean drop size and drop size distribution in an agitated vessel. Chem. Eng. Sci. 1999, 54
18
19 (19), 4211–4222.
20
21
22
(27) McManamey, W. J. Sauter mean and maximum drop diameters of liquid-liquid
23
24
25 dispersions in turbulent agitated vessels at low dispersed phase hold-up. Chem. Eng. Sci.
26
27 1979, 34 (3), 432–434.
28
29
30 (28) Wille, M.; Langer, G.; Werner, U. PDA Measurement of Drop Size Distribution for
31
32
33 Liquid-Liquid Dispersing in Agitated Tanks. Chem. Eng. Technol. 2001, 24 (5), 475–479.
34
35
36 (29) Biedermann, A.; Henzler, H.-J. Beanspruchung von Partikeln in Rührreaktoren (Stress on
37
38 Particles in Stirred-Tank Reactors). Chem. Ing. Tech. 1994, 66 (2), 209–211.
39
40
41
42
(30) Spicer, P. T.; Keller, W.; Pratsinis, S. E. The Effect of Impeller Type on Floc Size and
43
44 Structure during Shear-Induced Flocculation. J. Colloid Interface Sci. 1996, 184 (1), 112–
45
46 122.
47
48
49
(31) Zheng, J. L.; Shukla, V.; Wenger, K. S.; Fordyce, A. P.; Pedersen, A. G.; Marten, M. R.
50
51
52 Effects of Increased Impeller Power in a Production-Scale Aspergillus Oryzae
53
54 Fermentation. Biotechnol. Prog. 2002, 18 (3), 437–444.
55
56
57
58
59
60 ACS Paragon Plus Environment
43
Industrial & Engineering Chemistry Research Page 44 of 47

1
2
3 (32) Xia, J.-Y.; Wang, Y.-H.; Zhang, S.-L.; Chen, N.; Yin, P.; Zhuang, Y.-P.; Chu, J. Fluid
4
5
6 dynamics investigation of variant impeller combinations by simulation and fermentation
7
8 experiment. Biochem. Eng. J. 2009, 43 (3), 252–260.
9
10
11 (33) Hardy, N.; Augier, F.; Nienow, A. W.; Béal, C.; Ben Chaabane, F. Scale-up agitation
12
13
14
criteria for Trichoderma Reesei fermentation. Chem. Eng. Sci. 2017, 172, 158–168.
15
16
17 (34) Liepe, F.; Sperling, R.; Jembere, S. Rührwerke-Theoretische Grundlagen, Auslegung und
18
19 Bewertung; Fachhochschule Anhalt Köthen: Köthen, 1998.
20
21
22
(35) Wollny, S.; Sperling, R.; Kraume, M.; Ritter, J. Beanspruchung von Partikeln und
23
24
25 Fluidelementen beim Rühren (Stress of Particles and Fluid Elements during Stirring).
26
27 Chem. Ing. Tech. 2007, 79 (7), 1024–1028.
28
29
30 (36) Wollny, S. Experimentelle und Numerische Untersuchungen zur Partikelbeanspruchung in
31
32
33 Gerührten (Bio-)Reaktoren. Ph.D. Thesis, Technische Universität Berlin, 2010.
34
35
36 (37) Zhou, G.; Kresta, S. M. Evolution of drop size distribution in liquid-liquid dispersions for
37
38 various impellers. Chem. Eng. Sci. 1998, 53 (11), 2099–2113.
39
40
41
42
(38) Fries, T.; Dittler, I.; Blaschczok, K.; Löffelholz, C.; Dornfeld, W.; Schöb, R.; Drews, A.;
43
44 Eibl, D. Quantifizierung der hydromechanischen Beanspruchung von Pumpen auf
45
46 tierische Zellen mittels des nicht-biologischen Modellsystems Emulsion (Quantification of
47
48
the Mechanical Stress Exerted by Pumps on Mammalian Cells Using an Emulsion as a
49
50
51 Non-Biological Model System. Chem. Ing. Tech. 2016, 88 (1–2), 177–182.
52
53
54 (39) Ritter, J. Dispergierung und Phasentrennung in Gerührten Flüssig/flüssig-Systemen. Ph.D.
55
56
57
58
59
60 ACS Paragon Plus Environment
44
Page 45 of 47 Industrial & Engineering Chemistry Research

1
2
3 Thesis, Technische Universität Berlin, 2002.
4
5
6
7 (40) Peter, C. P.; Suzuki, Y.; Büchs, J. Hydromechanical Stress in Shake Flasks: Correlation
8
9 for the Maximum Local Energy Dissipation Rate. Biotechnol. Bioeng. 2006, 93 (6), 1164–
10
11 1176.
12
13
14
15
(41) Hoffmann, J.; Büscher, K.; Hempel, D. C. Ermittlung von Maximalen Scherspannungen in
16
17 Rührbehältern (Determination of Maximum Shear Stress in Stirred Tanks). Chem. Ing.
18
19 Tech. 1995, 67 (2), 210–214.
20
21
22
(42) Langer, G.; Deppe, A. Zum Verständnis der hydrodynamischen Beanspruchung von
23
24
25 Partikeln in turbulenten Rührerströmungen (On the Hydrodynamic Stress of Particles in
26
27 Turbulent Impeller Flows). Chem. Ing. Tech. 2000, 72 (1–2), 31–41.
28
29
30 (43) Stintzing, A.; Pilz, R. D.; Hempel, D. C.; Krull, R. Mechanische Beanspruchungen in
31
32
33 Mehrphasenreaktoren (Mechanical Stress in Multiphase Reactors). Chem. Ing. Tech. 2008,
34
35 80 (12), 1837–1842.
36
37
38 (44) Wollny, S.; Sperling, R. Partikelbeanspruchung in Gerührten Behältern (Particle Stress in
39
40
41
Stirred Tanks). Chem. Ing. Tech. 2007, 79 (3), 199–208.
42
43
44 (45) Pilz, R. D.; Hempel, D. C. Mechanical stress on suspended particles in two- and three-
45
46 phase airlift loop reactors and bubble columns. Chem. Eng. Sci. 2005, 60 (22), 6004–6012.
47
48
49
(46) Hinze, J. O. Fundamentals of the Hydrodynamic Mechanism of Splitting in Dispersion
50
51
52 Process. AIChE J. 1955, 1 (3), 289–295.
53
54
55 (47) Daub, A.; Böhm, M.; Delueg, S.; Mühlmann, M.; Schneider, G.; Büchs, J.
56
57
58
59
60 ACS Paragon Plus Environment
45
Industrial & Engineering Chemistry Research Page 46 of 47

1
2
3 Characterization of hydromechanical stress in aerated tanks up to 40 m3 scale by
4
5
6 measurement of maximum stable drop size. J. Biol. Eng. 2014, 8 (1), 1–14.
7
8
9 (48) Calabrese, R. V.; Wang, C. Y.; Bryner, N. P. Drop Breakup in Turbulent Stirred-Tank
10
11 Contactors Part III: Correlations for Mean Size and Drop Size Distribution. AIChE J.
12
13
14
1986, 32 (4), 677–681.
15
16
17 (49) Pacek, A. W.; Man, C. C.; Nienow, A. W. On the Sauter mean diameter and size
18
19 distributions in turbulent liquid/liquid dispersions in a stirred vessel. Chem. Eng. Sci.
20
21
1998, 53 (11), 2005–2011.
22
23
24
25 (50) Kraume, M.; Gäbler, A.; Schulze, K. Influence of Physical Properties on Drop Size
26
27 Distributions of Stirred Liquid-Liquid Dispersions. Chem. Eng. Technol. 2004, 27 (3),
28
29 330–334.
30
31
32
33 (51) Konno, M.; Kosaka, N.; Saito, S. Correlation of Transient Drop Sizes in Breakup Process
34
35 in Liquid-Liquid Agitation. J. Chem. Eng. Jpn 1993, 26 (1), 37–40.
36
37
38 (52) Chen, H. T.; Middleman, S. Drop Size Distribution in Agitated Liquid-Liquid Systems.
39
40
41
AIChE J. 1967, 13 (5), 989–995.
42
43
44 (53) Zhou, G.; Kresta, S. M. Impact of Tank Geometry on the Maximum Turbulence Energy
45
46 Dissipation Rate for Impellers. AIChE J. 1996, 42 (9), 2476–2490.
47
48
49
(54) Ranade, V. V.; Joshi, J. B. Flow Generated by Pitched Blade Turbines I: Measurements
50
51
52 Using Laser Doppler Anemometer. Chem. Eng. Commun. 1989, 81 (1), 197–224.
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
46
Page 47 of 47 Industrial & Engineering Chemistry Research

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

You might also like