You are on page 1of 147

INVESTIGATION OF DESIGN AND OPERATING PARAMETERS IN

PARTIALLY-FILLED RUBBER MIXING SIMULATIONS

A Thesis

Presented to

The Graduate Faculty of The University of Akron

In Partial Fulfillment

of the Requirements for the Degree

Master of Science

Suma Rani Das

October, 2016
INVESTIGATION OF DESIGN AND OPERATING PARAMETERS IN

PARTIALLY-FILLED RUBBER MIXING SIMULATIONS

Suma Rani Das

Thesis

Approved: Accepted:

Advisor Interim Dean of the College


Dr. Abhilash J. Chandy Dr. Donald P. Visco Jr.

Faculty Reader Dean of the Graduate School


Dr. Alex Povitsky Dr. Chand K. Midha

Faculty Reader Date


Dr. Jae-Won Choi

Department Chair
Dr. Sergio D. Felicelli

ii
ABSTRACT

The modern rubber industry is always in pursuit of improvements in the properties of the

final product resulting from the mixing of the rubber compounds with different fillers and

additives. Depending on the functional characteristics of the final product and thus the

compounding ingredients, different types of mixers can be used for the rubber mixing

process. Hence, the choice of an appropriate mixer is critical in achieving the proper dis-

tribution and dispersion of fillers in rubber, and a consistent product quality, as well as is

the attainment of high productivity. Besides rotor design, operational parameters such as

speed ratio and the orientation of the mixing rotors with respect to each other also play

significant role in the mixing performance. With the availability of high-performance com-

puting resources and high-fidelity computational fluid dynamics tools, understanding the

flow field and mixing characteristics associated with rotor orientations, speed ratios and

complex rotor geometries, has become more feasible over the last two decades. As part of

this effort, all the simulations here are carried out in a 75% fill chamber with two counter-

rotating rotors using a CFD code. In the phase angle and rotor design studies conducted

here, the rotors rotate at 20 rpm even speed, whereas for speed ratio study, only the left

rotor rotates at 20 rpm and the right rotor rotates at a speed, which is a multiple of 20 rpm

by the speed ratio specified. The computational models used in this research are based on a

iii
finite volume method to simulate a partially filled mixer equipped with different tangential

rotor types. The model solves for transient, isothermal and incompressible set of govern-

ing fluid equations for the mixing of non-Newtonian high-viscosity rubber. The research

here considers phase angles of 45◦ , 90◦ and 180◦ , speed ratios of 1.0, 1.125 and 1.5, and

rotor designs including 2-wing, 4-wing A and the 4-wing B rotors. Investigation of each

parameter type carried out separately.

The flow field is analyzed via pressure and velocity contours, mass flow patterns,

velocity vectors and particle trajectories. Dispersive mixing is evaluated through his-

tograms of mixing index, joint probability density functions of mixing index and shear rate,

and cumulative probability distribution functions of maximum shear stress experienced by

the particles. Distributive mixing is quantified statistically using cluster distribution index,

axial distribution, inter-chamber particles transfer, segregation scale and length of stretch.

The results helped in understanding the mixing process and material movement, thereby

generating information that could potentially improve the productivity and efficiency in tire

manufacturing process.

iv
ACKNOWLEDGEMENTS

First of all, the author would like to thank her adviser Dr. Abhilash J. Chandy for his

guidance, support, supervision, and insights throughout the research, as well as for showing

her right direction not just in research, but in other aspects of graduate life as well.

The author is grateful to Dr. Alex Povitsky and Dr. Jae-Won Choi for being in the

committee, and for their helpful comments and suggestions.

The author would also like to thank The Goodyear Tire & Rubber Company,

Akron, Ohio, for providing financial support.

Finally, the author would like to express her gratitude towards her parents, for

whom no amount of thanks is enough.

v
TABLE OF CONTENTS

Page

LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . viii

LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix

CHAPTER

I. INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

1.2 Investigation of rubber mixing . . . . . . . . . . . . . . . . . . . . . . . . 25

1.3 Objective . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

II. PROBLEM DESCRIPTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

2.1 Geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

2.2 Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

III. MATHEMATICAL FORMULATION . . . . . . . . . . . . . . . . . . . . . . . 52

IV. COMPUTATIONAL DETAILS . . . . . . . . . . . . . . . . . . . . . . . . . . 55

V. RESULTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

5.1 Mixing Measures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

5.2 Phase Angles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

5.3 Speed Ratios . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82

5.4 Rotor Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100

vi
5.5 Validation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118

VI. CONCLUSION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120

BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126

vii
LIST OF TABLES

Table Page

2.1 Rotor Geometry Specification . . . . . . . . . . . . . . . . . . . . . . . . . . 51

5.1 Percentage of Particles transferred at least once from one chamber to the other 95

viii
LIST OF FIGURES

Figure Page

1.1 Cross-section of an internal mixer [1]. . . . . . . . . . . . . . . . . . . . . . 3

1.2 Pointon’s machine for preparing rubber [2]. . . . . . . . . . . . . . . . . . . . 5

1.3 Banbury’s machine for treating rubber or heavy plastic materials [3]. . . . . . 8

1.4 Tyson and Comper’s four wing rubber mixer [4]. . . . . . . . . . . . . . . . . 8

1.5 Weidmann and Schmid’s mixing apparatus (a) Rotor geometry, (b) Ma-
terial motion [5]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.6 Nortey’s 2-wing non-intermeshing rotors (a) Rotor geometry, (b) Mate-
rial motion [6]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.7 Nortey’s 4-wing non-intermeshing rotors (a) Rotor geometry, (b) Mate-
rial motion [5]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.8 Takakura et al.’s 6-wing non-intermeshing rotors (a) Cross section of
rotor, (b) Material motion [7]. . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.9 Inoue et al.’s kneading apparatus (a) Rotor geometry, (b) Material motion [8]. 14

1.10 Inoue et al.’s batch mixer (a) Rotor geometry, (b) Material motion [9]. . . . . 14

1.11 Valmis et al.’s batch mixer (a) Rotor geometry, (b) Material motion [10]. . . . 16

1.12 Limper et al.’s batch mixer (a) Cross section of rotor geometry, (b) Ma-
terial motion [11]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
1.13 Yoshida et al.’s kneading rotor (a) Rotor geometry, (b) Material motion [12]. . 36

1.14 Cooke’s intermeshing rotors [13]. . . . . . . . . . . . . . . . . . . . . . . . . 37

ix
1.15 Wiedmann and Schmid’s intermeshing rotors [5]. . . . . . . . . . . . . . . . . 38

1.16 Johnson et al.’s interlocking mixing machine [14]. . . . . . . . . . . . . . . . 38

1.17 Passoni’s interlocking mixing machine [15]. . . . . . . . . . . . . . . . . . . 39

1.18 (a) GK-N series tangential mixers, (b) GK-E series intermeshing mixers,
(Courtesy of HF mixing group). . . . . . . . . . . . . . . . . . . . . . . . . . 40
1.19 Cooling systems used in the rotors (Courtesy of HF mixing group). . . . . . . 40

1.20 Ram movement [16]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

1.21 Typical mixing cycle [16]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

1.22 Distributive and dispersive mixing. . . . . . . . . . . . . . . . . . . . . . . . 42

2.1 2-wing rotor geometry: (a) Top view, (b) Cross-section at the middle. . . . . . 43

2.2 Orientation of the rotors for phase angle 45◦ (a), 90◦ (b) and 180◦ (c). . . . . . 45

2.3 Cross-section of the rotor along a plane in the middle at different time
instants during a rotor revolution, represented by different colors, for a
speed ratio of 1.125. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
2.4 Differential phase angles Vs. revolutions of the left (slower rotating) rotor . . 47

2.5 Geometry of the three rotors. . . . . . . . . . . . . . . . . . . . . . . . . . . 48

2.6 Comparison of average velocity of rubber between 100% and 75% fill
factors for 5 revolutions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
3.1 Change of non-Newtonian viscosity of rubber with shear rate based on
Carreau-Yasuda model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
4.1 2D unstructured quadrilateral mesh for phase angle 45◦ (a), 90◦ (b) and
180◦ (c). (d) Closer view of mesh at the tips. . . . . . . . . . . . . . . . . . . 55
4.2 Polyhedral mesh on the rotors. . . . . . . . . . . . . . . . . . . . . . . . . . . 59

4.3 Mesh on the cross-sections. . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

x
5.1 Pressure contours for phase angles 45◦ (left), 90◦ (middle) and 180◦
(right) at 84◦ (top) and 360◦ (bottom) rotations of the rotors after 19
revolutions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
5.2 Velocity contours for phase angles 45◦ (left), 90◦ (middle) and 180◦
(right) at 84◦ (top) and 360◦ (bottom) rotations of the rotors after 19
revolutions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
5.3 Histogram of mixing index calculated over last 15 revolutions for phase
angles 45◦ , 90◦ and 180◦ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
5.4 Cumulative probability distribution function of maximum shear stress
experienced by particles. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
5.5 (a) Initial position of the particle-clusters (magenta circular-middle, cyan
diamond-rotor tip gap, black triangle-left center near rotor wall), (b)
Particle-cluster dimension. . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
5.6 Particle distribution for phase angles 45◦ (left), 90◦ (middle) and 180◦
(right) after 5, 15 and 20 revolutions (from top to bottom). . . . . . . . . . . . 75
5.7 Evolution of cluster distribution index for the three phase angles and
three different initial cluster positions: (a) cluster-1 (b) cluster-2, and
(c) cluster-3. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
5.8 Evolution of cluster distribution index for all the three clusters combined. . . . 80

5.9 Transfer of particles from the (a) left to right chambers, and (b) right to
left chambers for three phase angles: 45◦ (green), 90◦ (green), and 180◦ (red). 81
5.10 Joint probability density function (jPDF) of the mixing index and shear
rate calculated over a period of 8 rotor revolutions of the left rotor after
12 revolutions for speed ratios of (a) 1, (b) 1.125 and (c) 1.5. . . . . . . . . . 85
5.11 Cumulative distribution of maximum shear stress experienced by the
particles over a period of one revolution . . . . . . . . . . . . . . . . . . . . 86
5.12 (a) Initial position of the particles (b) Probability distribution of the par-
ticle pair distances at the initial condition (red line); Also shown here
is the probability distribution of the particle pair distances of the ideal
distribution (black line with symbols). . . . . . . . . . . . . . . . . . . . . . 88
5.13 Particle distribution for speed ratios 1 (left), 1.125 (middle) and 1.5
(right) after 5, 15 and 20 revolutions (from top to bottom). . . . . . . . . . . . 91
5.14 Evolution of the cluster distribution index with time. . . . . . . . . . . . . . . 92

xi
5.15 Evolution of axial distribution with time. . . . . . . . . . . . . . . . . . . . . 93

5.16 Two types of particles with concentration 1 (green) and 0 (black) defined
at the end of 10 revolutions for scale of segregation calculation . . . . . . . . 94
5.17 Front view of the distribution of particles for speed ratios 1 (left), 1.125
(middle) and 1.5 (right) at 12 (top) and 20 (bottom) revolutions; The
green and black particles represent concentrations of 1 and 0, respectively. . . 94
5.18 Top view of the distribution of particles for speed ratios 1 (left), 1.125
(middle) and 1.5 (right) at 12 (top) and 20 (bottom) revolutions; The
green and black particles represent concentrations of 1 and 0, respectively. . . 97
5.19 Evolution of scale of segregation with time. . . . . . . . . . . . . . . . . . . . 98

5.20 Summary of (a) dispersive mixing and (b) distributive mixing capabili-
ties for three speed ratios: 1 (blue), 1.125 (red), and 1.5 (green). . . . . . . . . 99
5.21 Cross-sectional planes along the (a) axial, and (b) and (c) transverse di-
rections for analysis of flow rates (using (a) and (b)) and velocity vectors
(using (c)) at time, t=0. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
5.22 Instantaneous velocity vectors colored by velocity magnitude for 2-wing
(left), 4-wing A (middle) and 4-wing B (right) for cross-sections at 25%,
50% and 75% (from top to bottom) of the rotor length at 1.2 revolutions;
cross-sections shown in Figure 5.21c. . . . . . . . . . . . . . . . . . . . . . . 102
5.23 Instantaneous velocity vectors colored by velocity magnitude for 2-wing
(left), 4-wing A (middle) and 4-wing B (right) for cross-sections at 25%,
50% and 75% (from top to bottom) of the rotor length at 1.67 revolu-
tions; cross-sections shown in Figure 5.21c. . . . . . . . . . . . . . . . . . . 103
5.24 Instantaneous velocity vectors of 2-wing (left), 4-wing A (middle) and
4-wing B (right) rotor at 1.2 (top) and 1.67 revolutions (bottom), for the
cross section through the both rotor along their axis. . . . . . . . . . . . . . . 106
5.25 Evolution of normalized flow rate of 2-wing (left), 4-wing A (middle)
and 4-wing B (right) rotors for 2 revolutions from 15 to 17, through
the cross sections shown in Figure 5.21a; Lines are colored by cross-
sections shown in Figure 5.21a. . . . . . . . . . . . . . . . . . . . . . . . . . 107
5.26 Evolution of normalized flow rate of 2-wing (left), 4-wing A (middle)
and 4-wing B (right) rotors for 2 revolutions from 15 to 17, through
the cross sections shown in Figure 5.21b; Lines are colored by cross-
sections shown in Figure 5.21b. . . . . . . . . . . . . . . . . . . . . . . . . . 108

xii
5.27 Joint probability density of the mixing Index and shear rate calculated
over a period of one rotor revolution after 17 revolutions for the (a) 2-
wing (b) 4-wing A and (c) 4-wing B rotors. . . . . . . . . . . . . . . . . . . 110
5.28 Particle distribution for 2-wing (left), 4-wing A (middle) and 4-wing B
(right) at the initial condition, and after 1, 10 and 17 revolutions (from
top to bottom). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
5.29 Probability distribution of particles (a) at the initial time, and after 17
revolutions for (b) 2-wing, (c) 4-wing A and (d) 4-wing B rotors (solid
lines); Also shown are the corresponding ideal distribution of particles
for the three different rotors in (b), (c) and (d) (lines with symbols). . . . . . . 115
5.30 Evolution of the cluster distribution index with time. . . . . . . . . . . . . . . 116

5.31 Evolution of the mean length of stretch with time. . . . . . . . . . . . . . . . 116

5.32 Velocity profile along x-axis at the mid-cross section of the left half rotor
for 600k, 1.3M & 2M Cells. . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
5.33 Comparison of pressure profile obtained from Freakley and Patel’s ex-
periment at 50 rpm at the speed ratio of 1:1.125 [17] and simulation at
20 rpm even speed. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119

xiii
CHAPTER I

INTRODUCTION

Mixing is a basic and important step in the manufacturing process of polymer industries,

such as rubber compounding and plastics processing. The process of enhancing mechan-

ical, chemical and physical properties by reducing the non-uniformity of the ingredients,

known as mixing, affects the throughput as well as the final product quality [18]. Different

physical properties are attainable, depending on the range of concentration and type of the

ingredients. Tensile strength, fracture resistance, wearing and tearing resistance are some

examples of physical attributes in such a process. Besides polymer industries, high viscous

mixing has applications also in the field of food processing, paint mixing and so on.

Achieving uniformity is the primary goal of any mixing process. However, mixing

ingredients with highly viscous fluids can be very challenging and can affect productivity.

The way to obtain good mixing is to have a better understanding of the mixing dynamics,

material characteristics, effect of different process parameters on mixing and the choice of

an appropriate mixer. In addition, the efficiency of mixing depends on flow pattern, which

in turn depends on the geometry, as well as different operational parameters such as rotor

speed, speed ratio, as well as the orientation of the rotors [19].

1
1.1 Background

1.1.1 Rotor Design

Devices used for polymer mixing include single screw extruders, twin screw extruders, in-

ternal mixers, two-roll mills and continuous mixers. The purpose of any mixer is to reduce

non-uniformity while mixing one or more polymers with additives and fillers. Extruders

and continuous mixers are widely popular for plastic compounding [20]. In the rubber

industry, extruders and continuous mixers serve the purpose of finishing the mixing, re-

moving gaseous material, and defining shape after receiving the mixed compound from

internal mixer [20–22]. Two-roll mills are sometimes used for enhancing the dispersion of

fillers [18, 23]. In two-roll mills, it is possible to adjust the gap between the rolls to obtain

required dispersion. These devices are used for different applications, depending on the

mixing material and final product.

Internal batch mixers are primarily used for the sole purpose of rubber mixing

[22]. The history of using internal batch mixers for rubber mixing can be traced back

to the early 20th century [24]. With an increasing interest for rubber mixing in the tire

industry, internal batch mixers started to gain more popularity [24]. Internal batch mixers

comprise mainly of mixing rotors, a mixing chamber, a feeding hopper, a ram, a drop door,

a base plate and end plates (See Figure 1.1). Two C-shaped cylindrical halves enclose the

mixing rotors constituting the mixing chamber. The two ends are closed by the end plates.

Thus, the movements of materials are limited radially and axially. The feeding hopper

introduces the ingredients to be mixed inside the chamber. Inside the rectangular housing,

2
Figure 1.1: Cross-section of an internal mixer [1].

3
the ram can move up and down vertically. Lowering the ram pushes the materials down and

increases the pressure inside the chamber. It is possible to control the ram position further

in order to achieve the desired pressure. At its lowest position, the ram closes the mixing

chamber and the drop door is also kept closed to ensure that ingredients participate actively

in the mixing. The mixed material can be discharged through the drop door. Different

cooling and heating configurations are employed to maintain the desired temperature for

mixing. The most essential part of any mixing device is the set of mixing rotors. The

mixing rotors can be co-rotating or counter-rotating depending on the application. But in

an internal mixer, rotors are generally counter rotating. The effect of mixing depends on the

mixing rotor design. Mixing rotors are of basically two types: Non-intermeshing mixers

and intermeshing mixers.

Non-intermeshing or tangential mixers are placed parallel and do not interfere

with each other during mixing. Thus, there is always a constant gap between the two

rotors and it is possible to rotate each of them at different speeds. When the rotors run at

different speeds, there is a friction ratio, and when they run at equal speeds, the process

is referred to as even speed mixing. Usually the rotors are rotated counter-clock wise.

Various rotor types and profiles have been proposed and commercially utilized over the

years to improve mixing. The design of the rotors can vary depending on the number of

wings, wings length, shape of the wings, angles associated with wings, wing orientation

and their relative positions. Tangential rotors are well-known for its ease of feeding and

discharge [16].

In the beginning of the 20t h century, Kempeter, Pointon and Banbury filed patents

4
for tangential mixers separately, and today’s tangential mixers have similarities to those. In

1915, J. E. Pointon patented a rotor geometry particularly for mixing rubber (See Figure

1.2) [2]. The invention describes a rubber preparing machine that has a pair of intercom-

municating chambers, each containing one rotor. Each of these rotors has a spiral blade.

The spiral blade starts from one of the rotors, and ends before reaching the other end cre-

ating a passage. The rotors are arranged such that the passages from opposite rotors are

at the opposite ends. These passages help material transfer between the chambers. The

invention claims that such movement of material can be very effective in the mastication

and incorporation of rubber.

(a) (b)

Figure 1.2: Pointon’s machine for preparing rubber [2].

In 1916, Fernley H. Banbury patented a design of machine to treat rubber and

heavy plastics [3]. This invention is very similar to Pointon’s invention, but each rotor

has two blades, instead of the one in Pointon’s case. The blades on the rotors create acute

5
angles with stationary surfaces, and when rotated, presses the rubber against the surfaces.

The rotation of the blades pushes the rubber, and imparts rolling, kneading and squeezing

forces. Each rotor blade starts from the rotor ends, and terminates at the middle of the

rotor (See Figure 1.3. The blades move the materials from the rotor ends towards the

middle of the chamber applying a drag force. Banbury also suggested the use of a two-

cylinder chamber with opening at the middle for transfer of material. Material at the middle

mixes and exchanges between the rotors through the opening at the middle. According

to the patent, a weight (ram) can be applied on the upper part of the chamber to ensure

feeding and mixing of rubber between the rotors. Banbury filed another patent in 1916,

which introduced design of an automatic door and provides more detail on the ram [25].

According to the patent, lower surface of the chamber is replaced with a drop door. It is

possible to lock the door in tightly fit during the mixing period. The ram can be moved up

and down with a piston. The pressure within the chamber can be controlled using the ram

weight. A chute can be used to incorporate materials into the chamber. A water cooling

system was described as well, admitting water through hollow rotor blades and holes in the

chambers. The patent provided more details of rubber mixing machine compared to the

earlier designs.

The very first 4-wing rotor design was given by Lasch and Frei [24]. The 4-wing

rotor of Lasch and Frei was known as N-rotors, and was very popular in the 70’s [16].

Later, D. Z. Tyson and L. F. Comper from Goodyear Tire and Rubber company patented

a second 4-wing rotor design in 1966 [4]. Figure 1.4 shows the two rotors are parallel to

each other, but oriented in reversed direction. Each of these counter-rotating rotors has

6
two long wings at one end, and two short wings at the other end. The wings move the

materials back and forth along the axial direction of the rotors, as well as in transverse

direction. The patent claims to provide homogenously mixed rubber in a shorter amount

of time compared to the earlier inventions. According to the patent, the arrangement of the

wings and reduced cross section at the junction of the wings give better material intake and

can increase the efficiency between 25% and 50%.

Weidmann and Schmid from then Werner & Pfleiderer filed a patent for rotor

geometry of both intermeshing and non-intermeshing types [5]. The details of intermeshing

rotor have been discussed later in this section. In Weidmann et al.’s non-intermeshing

mixer, the rotors are placed parallel and mirror symmetric (See Figure 1.5). Principle

wings terminate before reaching the other end, and thus create passages. The short wings

at the two ends work as scraping tools, so that there is no stagnant material. The rotors

can have a cylindrical shaft core or conical core with aforementioned wing designs. The

conical cores will create a pressure gradient, not to mention the changing tip clearance

along the rotor axis. The spacious clearance will permit more material and give better

temperature distribution. Other recommended specifications are a maximum pitch angle of

40, a projected wing length to rotor length ratio of at least 0.5, and a clearance between

wing tip and chamber wall to chamber diameter ratio of at least 0.02. The patent claims

that the described design will have better heat transfer, and thus rotor can be operated at

higher speed.

Later around 1988, Nortey from Farrel Corporation along with his colleagues filed

a series of patents on both two wing and four wing rotor designs [1, 6, 26]. Nortey’s first

7
(a) (b)

Figure 1.3: Banbury’s machine for treating rubber or heavy plastic materials [3].

(a) (b)

Figure 1.4: Tyson and Comper’s four wing rubber mixer [4].

8
patent is on two wing rotors. Each rotor has two long wings, each coming from opposite

ends (See Figure 1.6). The rotors can be run at both equal and unequal speeds. The patent

provides experimental results for angles of 0◦ , 90◦ , 135◦ and 180◦ and suggests that for

optimum results a 180◦ phase angle should be used. The patent claims the design will

provide better mixing for first helix angle of 25◦ to 40◦ and second helix angle of 35◦ to

55◦ . Other specifications are rotor orientation of 176◦ to 184◦ , first twist angle of 80◦ to

110◦ and second twist angle of 90◦ to 120◦ . The argument is that the described design

provokes more wing to wing interaction, whereas in the prior designs, there is sitting of

material at the middle of the two rotors due to lack of interactions. The patent also specifies

its superiority in terms of higher fill factor capacity. The second patent of Nortey is for 4-

wing and 3-wing tangential rotors. Figure 1.7 shows each four wing rotor having one long

and one short wing originating from both ends. The patent describes increased twist angles

of wing, stimulating more homogenous mixing through rolling banks of material in axial

direction. Furthermore, the larger wing length ratio increases transfer of material from the

ends of the rotors. Patent claims more dispersion and thorough mixing is possible due to

additional four wings.

A 6-wing rotor designed by Harada et al. [27] had the issue of reduced capacity of

the rotor due to an increased number of wings. In an attempt to solve this problem, another

6-wing rotor was developed by Takakura from Kobe steel (Kabushiki Kaisha Kobe Seiko

Sho.) [7]. Each rotor has three equally spaced long wings starting from one end of the rotor

and three equally spaced short wings starting from the other end. The short wings scrape

off the material from the rotor end, preventing sticking of material (See Figure 1.8). The

9
patent argues that the 4-wing mixer does not give uniform mixing within the mixing time.

As a result mixing time gets longer, which may lead to vulcanization. Therefore, often, the

mixing is carried out in multiple stages of cooling and mixing while using 4-wing mixers.

But, Takakura’s 6-wing rotor has better heat transfer, and it is possible to reduce the number

of stages of mixing using this design.

Kimino Inoue and others from Kobe steel patented a rotor [8], which has at least

two long wings with a three-staged tip portion and three clearances in the axial direction

of each wing (See Figure 1.9). The design also consists of two short wings for scraping

off material at the rotor ends. The patent notes that, if the clearance between the cham-

ber wall and rotor tip decreases, shearing force increases and gives better dispersion. But

at higher rotor speeds, as tip clearance decreases, temperature of the mixing material in-

creases. Thus for a low temperature allowable material, the rotors speed would have to

be decreased, which leads to low productivity. As a result, for such material, rotors with

larger tip clearance is used. On the other hand, for high temperature allowable materials,

rotors with smaller tip clearance is used. Therefore, as the materials to be mixed change,

the rotors need to be changed as well, which patent describes as impractical. The objective

of this patent is to design a mixer with a plurality of the tip clearances that can be used in

various mixing conditions. Inoue et al. later provided a new mixing rotor design with plu-

rality of mixing rotors in 2010 [9]. The rotor consists of four long and four short wings (See

Figure 1.10). The long and short wings have different clearance sizes, and twists opposite

of the direction of rotation. The long wings follow a twist curve, such that twist angle be-

comes smaller towards the middle. The invention described claims to prevent diminishing

10
(a) (b)

Figure 1.5: Weidmann and Schmid’s mixing apparatus (a) Rotor geometry, (b) Material

motion [5].

(a) (b)

Figure 1.6: Nortey’s 2-wing non-intermeshing rotors (a) Rotor geometry, (b) Material mo-

tion [6].

11
of performance and productivity resulting from heat generation, and thereby increasing the

efficiency of the mixer.

At the beginning of the 21st century, Valsamis et al. from Farrel Corporation

developed a noteworthy rotor design shown in Figure 1.11 [10]. The patent specifies that,

performance of the earlier invented non-intermeshing mixers are limited by temperature

rise, and does not provide equally good dispersive and distributive mixing. The four wing

rotor described in the patent consists of two long wings and these long wings have different

helix and approach angles. Similarly, each rotor has two short wings, and these short wings

have different helix and approach angle. The designed rotor promotes more exchange and

interaction of materials between the two rotors and eliminates possible area of material

stagnation. A feature of this invention is that intensity of mixing can be varied by varying

rotor speed without resulting in unwanted temperature rises.

Andreas Limper along with his colleagues from HF mixing group offered a new

design on mixing rotor in 2009 [11]. Limper’s rotor has four long wings, which extend at

least over half of the axial length of the chamber (See Figure 1.12). Each wing has two

sections; first section has a helix angle greater than the second section. The wings divide

material to be mixed into four portions per rotor giving intensive dispersive and distributive

mixing.

A recent development on rubber mixing rotors is from Yoshida et el.’s (2015) of

Kobe Steel [12]. Yoshida et al.’s design includes two long wings and two short wings. The

short wings are linear and have length less than half of the rotor length. Each short wing

is placed behind each long wing in the rotational direction. The first long wing has a twist

12
(a) (b)

Figure 1.7: Nortey’s 4-wing non-intermeshing rotors (a) Rotor geometry, (b) Material mo-

tion [5].

(a) (b)

Figure 1.8: Takakura et al.’s 6-wing non-intermeshing rotors (a) Cross section of rotor, (b)

Material motion [7].

13
(a) (b)

Figure 1.9: Inoue et al.’s kneading apparatus (a) Rotor geometry, (b) Material motion [8].

(a) (b)

Figure 1.10: Inoue et al.’s batch mixer (a) Rotor geometry, (b) Material motion [9].

14
angle 15◦ to 35◦ and three different clearances, ranging from large to small in the axial

direction. The second long wing, on the other hand, is nonlinear, and has twist angle of 45◦

or more (See Figure 1.13).

Intermeshing mixers creates an intermeshing zone at the middle of the two rotors.

Maximum clearance between them is smaller than the sum of maximum radius of the two

rotors. So, it is necessary to rotate them always at constant and equal speed. Otherwise the

wings of the rotors will interfere and will be damaged.

The first intermeshing rotor in the rubber industry was developed by R. T. Cooke

in 1935 [13]. Cooke designed the rotor in an attempt to prevent overheating, reduce manual

labor, and increase productivity as well as to intensify the mixing process. Cooke’s design

has two rotors placed parallel to each other (See Figure 1.14). Each rotor has one long

spiral wing with a relatively wider peripheral face, compared to non-intermeshing rotors,

concentric with the rotor axis. The ends of the spiral are slightly tapered as they meet the

rotor ends. The long wing extends the total axial length of the rotor and occupies half of

its cylindrical side. The other half has two short wings at the two ends. The long wing of

the other rotor can move between these short wings.

Wiedmann and Schmid designed a intermeshing rotor in 1979 [5]. The rotors were

known as PES rotors [16] and later went through a lot of improvement over the decades.

The objective behind this invention was to increase material transport and obtain uniform

temperature distribution. Figure 1.15 shows the parallel rotors in intermesh with each

other. Each has one long wing at the middle and two wings at two ends of the rotor. Later,

upgraded versions of the design, such as PES 3, PES 5 and PES 6, attracted the attention

15
(a) (b)

Figure 1.11: Valmis et al.’s batch mixer (a) Rotor geometry, (b) Material motion [10].

(a) (b)

Figure 1.12: Limper et al.’s batch mixer (a) Cross section of rotor geometry, (b) Material

motion [11].

16
of tire industry.

Later in 1986, Johnson and his colleagues designed an intermeshing rotor based

on Cookes design [14]. Shown in Figure 1.16b, each rotor has one long wing with wider

wing tip. The wing starts from one end of the rotor and terminates before reaching the

other end creating a passage. There are two small wings placed before and after the long

wings at the two ends. During rotation of the counter-rotating rotor, one rotor’s long wing

gears up in between other rotor’s short wings. This can be seen as foot print in the Figure

1.16a.

In 1987, Passoni from Pomini Farrel, made a noteworthy contribution in the field

of rubber mixer by designing the first Variable Internal Clearance (VIC) intermeshing ro-

tors [15, 24]. Figure 1.17 shows Passoni’s internal mixer. The distance between the axes of

the rotors can be varied to achieve the best possible gap suitable for mixing process. The

rotors can be locked at desired gap, or the gap can vary during a mixing cycle.

Farrel Corporation, Pomini Plastics and Rubber, and Werener & Pfleiderer devel-

oped a lot of designs for rubber mixing over the years. As of 2010 Farrel Corporation,

Pomini Plastics and Rubber, and Werner & Pfleiderer became wholly-owned subsidiaries

of HF mixing group. Some standard products of HF mixing group are the NST tangential

rotor, NR5 intermeshing rotor, GK-N series tangential rotor, and GK-E series intermeshing

rotor. According to the HF mixing group, NST rotor gives perfect combination of material

flow and shear intensity, while NR5 rotor provides high thermal efficiency. The features

of GK-N type rotors (Figure 1.18a) are high fill factor, and less feeding and discharging

time. On the other hand GK-E type intermeshing rotor (Figure 1.18b) is good for tem-

17
perature sensitive compounds and can reduce mixing time and cost tremendously. Other

related products from Kobe steel include 2-wing S tangential rotor, 4-wing H tangential

rotor, 4-wing N tangential rotor, 6-wing VCMT tangential rotor and intermeshing rotor.

A comparatively newer concept in rubber mixing is tandem mixing. Tandem mix-

ing system uses a combination of a conventional internal mixer with ram and a larger

mixer without any ram. In the conventional internal mixer filler dispersion is done under

ram pressure. This stage might in fact increase temperature. The compound mixed in this

stage is then discharged to tandem mixer for the next stage of mixing. Larger surface area

of the tandem mixer cools down the compound and continues further mixing. Reactive

substances are added to the compound in the tandem mixer. In this way, mixing happens

continuously in a tandem mixer and reduces the time of stopping and changing of mixers

involved in a multistage mixing.

1.1.2 Material properties

During mixing modeling, while choosing rotor design, it is also essential to understand the

characteristics of the raw materials. Rubber is an elastomer that allows large deformation

and 100% recovery of the deformation without breaking. However, under high stress and

low temperature, rubber can transform into a glassy substance [28]. At high temperature,

rubber acts like viscous fluid. Mixing is done at elevated temperatures, when rubber acts

like a viscous fluid. Although for better dispersion, milling can be done at lower tempera-

ture before mixing [28]. Depending on the application, such as sealing, coating, insulating,

or tire making, different types of rubber or a combination of them can be used. Mostly

18
used rubbers are Natural Rubber (NR), Styrene Butadiene Rubber (SBR), EPDM, Buta-

diene Rubber (BR) and so on. NR is very viscous and has high molecular weight. It is

temperature sensitive, but at low temperature, it has tensile strength and good resistance

to tearing and abrasion. NR is largely used for sealing and vibration damping. SBR is

extensively used in tire industry for its lower temperature sensitivity and high wear, tear

and abrasion resistance. EPDM is mainly a polymer blend of ethylene and propylene. Due

to it’s excellent resistance to heat, chemical and water, it is used for sealing hot water and

steam in faucets. BR is sometimes used in blend with SBR in tire industry. Viscosity

of BR does not change much with temperature, and the sticky nature of BR helps absorb

fillers and oils. Other types of rubber includes Nitrile Butadiene Rubber (NBR), Butyl

Rubber, Fluor Rubber, Chloroprene Rubber (CR), Polyethylacrylate etc. NBR has high

viscosity and good resistance to abrasion; Butyl Rubber is less viscous but very reactive;

Fluor Rubber is highly viscous, resistant to chemicals even at higher temperature; and CR

is very adhesive and corrosive, but gives good dispersion. Masterbatches are another kind

of rubber, which has oil or carbon black blended with them. The ingredients such as, car-

bon black, silica, antioxidant, oil, resins etc., when mixed with rubber contributes to the

improvement of its properties. For instance, carbon black is used as reinforcing filler. De-

pending on which grade of carbon black is being used, it can improve hardness, tensile

strength, tearing and wearing resistance, electrical conductivity and so on. Introducing sil-

ica in the mixing of rubber, reduced rolling resistance by 25%, which in turn reduced fuel

consumption by 5% [16]. Oil is added to the rubber to increase plasticity and durability.

Plasticizer, oil or other liquid additives are added during mixing to prevent glass formation,

19
as well as to provide desired properties to final product. Resins are added to the rubber in

tire industry to increase tackiness. Chemicals such as sulphur, zinc oxide are added as part

of the curing process. All these ingredients can have desirable effect on rubber properties,

if mixed properly.

1.1.3 Process Parameters

In order to bring out the best of ingredient properties, it is crucial to understand, observe

and control the process parameters such as: fill factor, speed, friction speed ratio, phase

angle, ram pressure, and temperature.

Fill factor is one of the most important influential parameter on mixing quality and

productivity. Fill factor can be defined as the ratio of volume of chamber space occupied by

rubber to the total volume available to be occupied. Higher filler factor does not necessarily

lead to higher productivity. It may require more time to achieve proper mixing with a higher

fill factor. It would be quite difficult and power consuming to run the rotors at higher speed

at higher fill factor. Besides, high fill factor leads to problems in terms of controlling the

temperature, due to viscous heating. Overfilling may cause some rubber to stick to the

discharge door and chamber throat, and not take part in mixing at all. On the other hand,

too much under-filling is uneconomical. If the amount of rubber is too small, there will be

little interaction between rubber and filler, leading to poor mixing. Therefore, search for

an optimum fill factor is inevitable. With an optimum fill-factor, rubber has enough space

to move around and provide effective mixing through folding and stretching. Commonly

used fill factor falls in between 0.65-0.85, depending on the material properties and rotor

20
type [21]. Fill factor tends to be lower for highly viscous material, whereas it is possible

to use a higher fill factor for low viscous material. Optimum fill factor also gives better

heat transfer from rubber to the cooling surfaces through both conduction and convection.

Rubber is a poor conductor, so conductivity plays very little role in heat transfer. So, the

heat transfer largely depends on the convection, which is controlled by fill factor.

Another significant factor of consideration during mixing is temperature. Mixing

happens in multiple stages. It is essential to control temperature in all the stages, especially

the stages where chemical reaction happens. Besides, temperature can markedly affect

dispersive and distributive mixing of fillers. Dispersive mixing is process of breaking the

fillers into small pieces, and distributive mixing is the process of spreading the fillers homo-

geneously in the entire domain. At high temperature, the viscosity reduces and increases

mobility of material, but reduces stress on the fillers. So, higher temperature can help dis-

tributive mixing and hinder dispersive mixing. Moreover, excessive temperature can cause

damage to the rotor surfaces. Temperature inside the chamber depends on the heat transfer

rate from the rubber to the chamber and the rotor walls. Temperature can be a limiting

factor and can create many issues especially, at high speed mixing. All the surfaces avail-

able are used to cool down overheated material from mixing. Typically, water is circulated

over the chamber wall, ram, discharge door, and through rotors [21]. Water can be sprayed

over the chamber wall and rotor wall or passed through drilled passage. Figure 1.19 shows

two cooling system from Farrel Corporation. Relatively, newer inventions use steam heat-

ing to deal with the condensation on chamber and rotor surfaces at cold environment [21].

Cooling and heating are controlled using electrical control system.

21
Optimum rotor speed depends on the design of the rotors as well as the material

to be mixed. In order to maximize the dispersive mixing, the material’s high viscosity

is maintained by running the rotors at lower rpm. At high rpm there is an increase in

temperature, which in turn, reduces the viscosity. Besides viscous heating, viscosity can

also reduce at high shear imparted by the high rotor speed. But, running the rotors at

lower rpm takes longer time to mix. Distributive mixing on the other hand, improves

with increasing rotor speed. Due to this conflicting effect of rotor speed on dispersive and

distributive mixing, the whole mixing process is carried out in multiple stages.

Both of the mixing rotors can be rotated at either same or different speeds. When

two rotors are rotated at the same speed it is known as even speed mixing. When two rotors

run at different speed, their speed ratio is known as friction speed ratio. The difference

in speed creates a velocity gradient between the two chamber lobes that contain the rotors.

This velocity gradient can be very effective in dispersive and distributive mixing depending

on the rotor design.

Furthermore, most of the times, rotors are run at phase angle with each other. The

difference in the angular position between the two rotors facilitates better movement of the

material resulting in better distribution of the fillers. Depending on the rotor design the

phase angle can vary.

Higher ram pressure is desirable for proper dispersion of fillers. Substantially

low ram pressure can result in very poor mixing. Moreover, the ram forces the rubber

and ingredient to engage in mixing. The pressure applied by the ram should be sufficient,

so that the pressure inside the chamber cannot cause upraising of the material. A regular

22
periodic upraising of material can happen when the rotors tips come together and push the

materials. Figure 1.20 shows the cyclic movement of ram to maintain its position.

Mixing of rubber with fillers, chemicals or liquid additives happen in four different

stages: mastication, incorporation, dispersion and distribution (See Figure 1.21). Mastica-

tion decreases viscosity of the polymer by reducing the chain length. In the incorporation

process, fillers and additives get soaked into rubber. Mastication and incorporation gener-

ally happen in the early stages of mixing. The topics of interest in this study are mostly

related to homogenization through distributive and dispersive mixing.

1.1.4 Mixing characteristics

Mixing characteristics is determined based on both dispersion and distribution of fillers

[29]. The complex geometry of rotors, their relative orientation and speed cause a variation

in local forward, backward and transverse flow patterns, and are critical in analyzing dis-

persive and distributive mixing indexes [30]. Fillers, such as silica and carbon black that

are typically mixed into rubber during the manufacturing process are difficult to disperse,

and thus dispersion efficiency has a significant influence on the final product [31]. So a

better understanding of dispersion along with distribution is necessary.

Dispersion is the process of breaking down the ingredients such as solid agglomer-

ate or liquid droplets, held together by cohesive forces, whereas distribution is the process

of spreading broken agglomerate throughout the fluid domain [32] (See Figure 1.22). Dis-

persive mixing is achieved by prevailing shear and elongational forces in the mixer. While

shear stress is essential in breaking up the (solid and liquid) agglomerates, elongational

23
components characterize the flow strength. Studies have shown that elongation has more

effect on agglomerate dispersion than simple shear flow [33]. In a complex geometry how-

ever, as is the case here, there is a superposition of flow regimes ranging from pure rotation

to pure elongation. This makes it necessary to determine elongation along with shear rate

for assessing dispersive mixing. There are lot of experimental techniques available such

as optical microscopy, electrical resistivity, optical profilometry, and electron microscopy

for dispersion measurements, which have their own limitations either in terms of cost or

accuracy [34]. One way to quantify this characteristic in numerical simulations is to cal-

culate the mixing index [35]. Even though challenges still remain in the field to find a

universal statistical quantity that can fully explain the dispersive mixing, quantities such as

mixing index and maximum shear stress, which are presented here, will give a reasonable

description of the mechanism.

Furthermore, while the breakup of the agglomerate into fine particles is important,

the effectiveness of the mixer can also be evaluated by the homogeneity of the distribution

of the small particles throughout the domain, including the stretching of the agglomerate,

i.e. the distributive mixing [36]. In order to characterize distributive mixing, a large set

of massless particles can be used. Qualitatively, distribution can be analyzed from a spatial

distribution of fictitious particles, but more specifically they can be quantitatively assessed

using parameters such as segregation scale, cluster distribution index, axial distribution and

particle transfers between chambers [37–39]. In addition, stretching can be determined

using a length of stretch [36]. This study uses all these parameters in evaluating mixing

efficiency for various operational conditions and rotor designs.

24
1.2 Investigation of rubber mixing

Several studies have been conducted both experimentally and numerically in the past, in

order to understand effect of different parameters on the flow field and mixing efficiency.

For instance, Freakley and coworkers made a noteworthy contribution in the field of rubber

mixing by developing an experimental setup to visualize rubber mixing process. Freakley

and Idris proposed to replace the front and side of mixing chamber with glass, so that move-

ment of the material could be analyzed [40]. They argued that, even though measurement

of several parameters was helpful, flow visualization helped in troubleshooting of prob-

lems. Along with flow visualization, pressure was also obtained from experiments, and

stress was determined using a mathematical model to compare fill factors. Higher pressure

was detected for higher fill factors. Freakley along with Patel later performed a detailed

analysis of flow and mixing characteristics for different fill factors with instrumented BR

Banbury and biconical rheometer [17]. Mathematical model was also developed in an at-

tempt to determine velocity, stress, and power distribution. The authors found out that at

high rotor speeds and low fill factors distributive mixing was poor. Later, they also stud-

ied the pressure and temperature developed in rubber mixing with intermeshing rotors [41].

Two rotor designs were studied at selected fill factors and rotor speeds. The authors found a

high shear stress distribution at the clearance between rotor tip and chamber wall, and also

in the rotor channels. The authors also discovered material to be viscous in the channels

and viscoelastic at the clearances.

James L. White and coworkers conducted several experiments to understand mix-

25
ing of rubber. Tokita and White studied behavior of gum elastomer in a two-roll mill at

different temperature [23]. The study showed that at high temperature, elastomers behave

like polymer solution and at low temperatures, they can crumble into powder. Also math-

ematical analysis was carried out to compute stress and velocity. In the following years,

White and coworkers did many experimental studies to visualize and understand flow of

rubber and other elastomers mixing with fillers and additives. Freakley et al.’s experimen-

tal setup was adopted by Min and White to visualize motion of various elastomers and

molten plastics in Banbury mixer [42]. SBR and BR found to be tearing when stretched,

causing stagnant regions below the ram. NR does not tear when stretched, but rather bands

on the rotor circumference. Addition of carbon black and oil with different type of elas-

tomers such as BR, NR, and SBR using various rotor designs was investigated by Min and

White [43]. Torque, temperature and power consumption were also analyzed. From obser-

vations of flow of rubber and ingredients, authors were able to detect stretching, tearing and

stagnant regimes. Later, Min used a flow visualization experimental setup to study blend-

ing of ethylene-propyrene terpolymer (EPDM) and a thermoplastic, polystyrene (PS) [44].

Phase morphology of the EPDM/PS blend was investigated using a scanning electron mi-

croscope. At low temperature, EPDM exhibits brittle nature and PS forms bands on the

chamber wall. At high temperature, EPDM forms sheets and PS exhibits liquid like flow.

Morikawa et al. studied flow visualization of gum rubber compounding with carbon black

and oil [45]. According to their research, better homogenization with carbon black is ob-

tained if rubber is pre-masticated, and oil is added at the later stage of mixing. Later, the

authors investigated compounding of blends of elastomers with carbon black and oil [46].

26
Color pigment was added for homogenization study. Authors detected improvements in

mixing, when carbon black and oil were added to blends of elastomers. Other work from

Kim and White include study of the effect of different fill factors by measuring torque,

pressure, number of circulations per fill factor, and flow field analysis [47]. Koolhiran and

White compared intermeshing and tangential rotors in dispersing carbon black, and found

intermeshing rotors gave better dispersion compared to tangential rotors.

Kawanishi and coworker also did some extensive experimental flow analysis on

rubber mixing, by changing different parameters of rotor design [48]. Motion of pig-

mented silicone was observed using a glass window at the front of the chamber. It was

demonstrated that, the homogenization time became shorter with the increase of rotor tip

clearance, overlap ratio of the wings and twist angle of wings, and decrease of ratio of rotor

speeds. Influence of rotor designs and mixing conditions on physical properties of rubber

compound was studied to predict the relationship that existed between them [49]. Most of

the flow analysis experiments were done using a scale-down laboratory mixer. So, Kawan-

ishi et al. investigated parameters of rotor design and mixing conditions to reproduce the

compound from the large scale industrial mixer using a laboratory scale mixer [50].

Other studies include Cottons methods development to measure dispersibility of

carbon blacks experimentally against different operating conditions and material proper-

ties [51]. Griffith et al. studied flow of natural rubber using a Brabender mixer with

Banbury type rotor [52]. Color was used in rubber to study the effect of process parameters

and randomness in the flow [53]. The dependence of mixing energy on different types of

fillers was studied experimentally by Leblanc and Lionnet [54]. Toh et al. experimentally

27
studied the effect of twist direction of rotors in the mixing of butadiene rubber with ZnO,

using three kinds of rotors [55]. Kaewsakul et al. experimented to optimize mixing tem-

perature and mixing interval of silica-silane-rubber [56]. Ghari et al. did experiments to

compare single phase and dual phase fillers in natural rubber as well as investigated effect

of sequencing of dual phase fillers [57]. Coming across all these researches, the importance

to study flow of rubber, as well as dispersive and distributive mixing is fittingly realized.

Although these experiments do provide invaluable real data, they are expensive

and difficult to set up. Besides, it is not possible to replicate the exact industrial situation in

the experimental setup and obtain useful data from it. It is almost impossible to determine

dispersive and distributive mixing capability from the experimental data directly. On the

other hand, numerical studies can provide statistical tools that can help understand the

effect of parameters more rigorously, without disturbing the flow field, which in turn can

assist in narrowing down the choices for experiments. Due to recent improvements in

computing techniques and resources, modeling of flow and the associated mixing processes

are capable of providing comprehensive data that can help optimize parameters and rotor

designs used in mixing.

Over the last few decades there have been several studies featuring 2D and 3D

numerical simulations of polymer processing and rubber compounds mixing in fully-filled

chambers including extruders and roll mills that employed both in-house and commercial

CFD codes. For instance, the need for a numerical model in roll milling was realized by

Kiparissides and Vlachopoulos in 1976 [58]. The authors used the Galerkin method to

perform finite element analysis of Newtonian and non-Newtonian flow at the nip region

28
of roll mills. Non-Newtonian behavior of fluid was defined by both power-law and hyper-

bolic tangent model. The lubrication approximation was used to determine pressure for

both symmetric and asymmetric calendering. In spite of the limitations related to the 2D

approximation, great effort was offered in the field of numerical analysis of rubber mixing.

Along with experimental studies of rubber mixing, White and coworkers con-

ducted some numerical analysis as well. Hydrodynamic lubrication theory was used to

stimulate Newtonian flow in internal mixer and modular corotating intermeshing twin-

screw extruders [59]. Pressure profile and flow field was determined in this isothermal

fully filled simulation. Kim et al. also attempted to simulate mixing of Newtonian fluid

using two counter rotating non-intermeshing rotors in order to obtain global circulation pat-

tern [60]. Kim and White later conducted non-isothermal simulations of non-Newtonian

flow using 3D internal mixer [61]. The limitation here was the use of unwound geometry,

which neglects the curvature of the rotors, which in turn might affect the conclusions on

mixing.

Yagii and Kawanishi noted that, it is not possible to measure flow pattern, shear

strain and stress, and pressure experimentally [62]. Flow visualization through transparent

walls is not perfect either. So, they decided to model mixing of rubber using finite element

method. The chamber was assumed to be fully-filled, and the material was assumed incom-

pressible and a power law liquid. The model helped in better understanding of dispersive

and distributive mixing. The model developed was 2D, and hence did have the limitation

of neglecting axial flow.

Manas-Zloczower has made significant impact in the computational study of rub-

29
ber mixing. Manas-Zloczower and Feke studied dispersive mixing of the agglomerates

in four different flow fields: simple shear flow, pure elongational flow, uniaxial flow and

biaxial flow [63]. The study found pure elongation to be more efficient in agglomerate

breakup than simple shear, biaxial extension being the most efficient. Cheng and Manas-

Zloczower studied the velocity and pressure profile of a 2D Banbury mixer using the fluid

dynamics analysis package, FIDAP, that implements a finite element method [64]. They

also studied effects of design and processing variables such as rotor tip gap, rotor speed and

speed ratio for a Banbury mixer calculating mixing index, shear rate and shear stress [65].

Mixing index, a parameter introduced by the authors, is now widely used for assessing dis-

persive mixing. The authors’ key contribution was the development of parameters such as

flow strength, mixing index and length of stretch for rubber mixing to quantify dispersion

and distribution characteristics [35, 39]. Later, the authors looked into the effects of rota-

tional speed on the distributive mixing, by calculating length of stretch and area of stretch,

and concluded that higher rotational speed increases distributive mixing efficiency [38].

Wong and Manas-Zloczower studied three different phase angles along with uneven speed

cases to determine optimum operating modes [37]. They used particle trajectory snapshots

for qualitative assessment of the distributive mixing. Calculation of pairwise correlation

functions for particles, and from that cluster distribution indices was introduced for a quan-

titative assessment of distributive mixing. For uneven speed cases, authors found speed

ratio close to 1 to be give the best mixing. Yang and Manas-Zloczower explored the ef-

fect of a changing gap on distributive and dispersive mixing performance of the Variable

Intermeshing Clearance (VIC) mixer by calculating cluster distribution index, mixing in-

30
dex, shear rate, and also observing pressure and velocity profiles of the mixer [33]. Yao

and Manas-Zloczower studied two different design of VIC in comparison to original de-

sign [66]. Dispersive mixing was evaluated in terms of shear stress and elongational flow

components, whereas distributive mixing was evaluated in terms of particle trajectory snap-

shots and cluster distribution index. VIC with enlarged chamber and smaller clearance

showed better dispersion and distribution. Authors also investigated distributive mixing in

the axial direction for four different designs of continuous mixer [67]. More studies from

Manas-Zloczower along with others, include modeling of the batch mixer with a simplified

double-Coutte flow geometry in the laminar and turbulent flow regimes [68], development

of mixing index to study dispersive and distributive mixing in scale-up twin-flight sin-

gle screw extruder [32], and development of method using Renyi entropy to characterize

distributive mixing in a twin-flight single screw extruder as well as to examine different

processing conditions [69, 70]. However all of these studies were based on the assumption

that the mixing chamber was fully filled in contrast to reality where chamber is partially

filled.

Yao and Manas-Zloczower also studied the shape of the free surface at the bank

and nip region of the two roll-mill using finite element method. The shape of the in-

terface was guessed initially and then changed at each iteration until convergence was

achieved [71]. The disadvantage of using a 2D model is that axial flow and distribution

are completely neglected, which in turn might affect the conclusions with regard to dis-

tributive and dispersive mixing characteristics.

A slightly different approach was adopted by Gramann and Osswald [72]. The

31
boundary element method was used to simulate polymer mixing process in Banbury type

mixer. The boundary element method eliminates the cumbersomeness associated with

meshing and re-meshing and gives more accuracy for higher order derivatives. Hutchinson,

Rios and Osswald later used the same technique to develop a new mixing index and to com-

pare performance for different rotor designs and mixing conditions [73]. Rauwendaal et al.

also used the boundary element method to investigate dispersive mixing in extruders and

internal mixer. However, these simulations ignore the axial distribution and non-Newtonian

effect of polymers.

Furthermore, the slip-stick boundary condition was analyzed in the study by Al-

steens et al. [74], and a parametric study of twin screw extruder was conducted as well [75].

Dispersive mixing of agglomerates using the mass density function was characterized by

Collin et al. [31]. Salahudeen et al. performed a comparative study of three different rotor

designs: cam, banbury and roller, through observations of velocity profiles and particle

trajectories [76]. Simulations were conducted using the FEM solver Polyflow and mesh

superposition technique was used for rotational motion of the rotors. The Carreau-Yasuda

model was used to represent the rubber viscosity, and mixing performance of the rotors

was quantified using length of stretch, instantaneous efficiency and time average efficiency.

Roller rotors were found to be more effective compared to the other two. Salahudeen and

others also studied the optimization of rotor speed using cam rotors based on stretching,

efficiency and viscous heating characteristics. Authors found that while viscous heating

and length of stretch increased with rotor speed, the efficiency decreased [77].

High viscous fluid mixing and extrusion is also of interest in the food industry.

32
Dhanasekharan and Kokini carried out numerical simulations of wheat dough extrusion

using a commercial finite element package known as Polyflow [78]. Effect of scaling was

explored by varying geometric parameters of single screw extruders. Connelly and Kokini

analyzed the effects of rheology with different viscoelastic and shear thinning models, and

evaluated the effectiveness of mixing using statistical quantities such as segregation scale,

cluster distribution index, mixing index, length of stretch and efficiency of mixing [79].

They also examined the ability of single and twin screw extruders for dough mixing using

finite element methods [36]. More recently, the effect of the change in paddle stagger

angle on the velocity distribution of the corn syrup was studied by Vyakaranam et al. [30],

and the effects of mixer speed, flow rate and paddle angle for mixing of wheat flour were

investigated using velocity vectors, shear rate and mixing index by Rathod et al. [80].

Nassehi and Salemi also developed a finite element model based on Galerkin

method to simulate non-isothermal viscometric flows in rubber mixing [81]. The Power-

low method was used for determining the non-Newtonian viscosity of rubber. Flow pattern,

temperature and pressure profiles were obtained and compared for different blade geome-

tries. Their model provided insights on necessity and validity of using numerical method.

Much of the published literature dealing with the mixing processes of polymer as-

sume a fully-filled chamber, in order to avoid the complexity associated with the calculation

of free surface, when in reality the chamber is never fully-filled. There are a few studies

on partially filled chambers, including Nassehi and Ghoreish’s free surface simulation us-

ing volume of fluid (VOF) method in Eulerian framework [82]. Fluid in the mixer was

considered as incompressible and air phase was considered compressible. They were able

33
to successfully predict the flow pattern in transient non-isothermal, non-Newtonian flow

simulation. Another partially-filled study done by Nassehi and Ghoreishy [83] was a finite

element simulation of internal mixer with long blade tips, where a Eulerian-Lagrangian

method for free surfaces was used in order to solve mixing of non-Newtonian fluid with

intermeshing type mixers in a 2D model. But these simulations ignored axial distribution,

which can be critical, especially in distributive mixing.

Advancement in the field of CFD, enabled Pokriefke to simulate viscous flow in a

partially-filled co-rotating twin screw extruder and compare with the fully-filled case [84].

The 3D simulation model used a finite volume and Eulerian multiphase method, while the

free surface flow was solved using the local volume fraction of a cell. One of the more

recent studies on a partially-filled internal batch mixer was a CFD study of the distributive

mixing characteristics for two different geometries and two different fill factors [85], where

a volume of fluid (VOF) method was used to calculate the free surface flow.

1.3 Objective

The objective of the current study is to develop a computational methodology to optimize

and analyze the effect of different rotor designs and operational conditions such as phase

angle and speed ratio through parametric studies of batch mixer in a chamber partially filled

with rubber. Phase angle is the angle at which the two rotors are oriented with respect to

each other during mixing. Since there has never been a computational investigation of

phase angle in partially filled chamber, here the effort was made to offer a better judgment

through two dimensional incompressible isothermal two phase (rubber+air) simulation.


34
The two rotors running at different speeds, where rotor-to-rotor orientation , i.e. phase an-

gle, is different at the different instances of a revolution can potentially improve dispersive

mixing by enhancing the elongational flow as well as distributive mixing by introducing

more randomization [20]. To the author’s knowledge, there have not been any 3D stud-

ies investigating the effect of different speed ratios under partially filled conditions. The

current study attempts to demonstrate this phenomena by considering two different speed

ratios: 1.125 and 1.5, in comparison with even speed (i.e. speed ratio of 1.0) in terms of dis-

persive and distributive mixing. Finally, it is also objective of this research to assess three

different rotor designs, the 2-wing, 4-wing A and 4-wing B rotor, in terms of its overall dis-

persive and distributive mixing characteristics for partially filled chamber considering the

fact that length, angle and number of the wings, and also their position with respect to each

other, control the material movement in the axial and transverse directions and randomizes

the blending of material to provide improved performance and productivity.

35
(a) (b)

Figure 1.13: Yoshida et al.’s kneading rotor (a) Rotor geometry, (b) Material motion [12].

36
(a) (b)

Figure 1.14: Cooke’s intermeshing rotors [13].

37
(a) (b)

Figure 1.15: Wiedmann and Schmid’s intermeshing rotors [5].

(a) (b)

Figure 1.16: Johnson et al.’s interlocking mixing machine [14].

38
(a) (b)

Figure 1.17: Passoni’s interlocking mixing machine [15].

39
(a) (b)

Figure 1.18: (a) GK-N series tangential mixers, (b) GK-E series intermeshing mixers,

(Courtesy of HF mixing group).

Figure 1.19: Cooling systems used in the rotors (Courtesy of HF mixing group).

40
Figure 1.20: Ram movement [16].

Figure 1.21: Typical mixing cycle [16].

41
Figure 1.22: Distributive and dispersive mixing.

42
CHAPTER II

PROBLEM DESCRIPTION

2.1 Geometry

As a part of the current research effort, phase angle and friction speed ratio were studied

for two wing rotors. Three different designs of the mixing rotors, rotating even speed, were

also studied.

2.1.1 Phase Angle

(a) (b)

Figure 2.1: 2-wing rotor geometry: (a) Top view, (b) Cross-section at the middle.

The orientation of rotors with respect to each other plays a big role in the mix-

ing of rubber. Here, three different rotor orientations will be studied in a partially filled
43
chamber. The geometry used here is a two-dimensional (2D) mid-cross section of a three-

dimensional (3D) model, built following the US patent 4,714,350 [6] (See Figure 2.1).

Detailed specification of the 3D geometry of the 2-wing rotor can be found in Table 2.1.

The rotors in the figure 2.1 are oriented at 180◦ with respect to each other. The other two

orientations being studied are 45◦ and 90◦ (See figure 2.2). While in operation, the left

rotor rotates clockwise, and the right rotor rotates counter-clockwise, about their own axes.

Each of these rotors has two helical long wings originating from opposite ends. These

kinds of rotors are very common in the mixing of polymer materials such as rubber and

plastics with their reinforcements. Frequently-used phase angles have been taken under

consideration in the current study [30, 37].

2.1.2 Speed Ratio

In order to study the mixing effect of different speed ratios, a 3D geometry of a 2-wing

non-intermeshing rotor, similar to the rotors of phase angle study, has been employed here.

The rotors are shown in Figure 2.1 and are oriented at 180◦ with each other. The left and

right rotors rotate clockwise and counter-clockwise about their axes, respectively. Each of

the rotor has two helical long wings originating from opposite ends. These kinds of rotors

are widely used in batch mixing of polymer materials such as rubber and plastics. The

rotor design used here follows the US patent 4,714,350 [6]. The specifications are listed

in Table 2.1. See [6] for a description of these parameters. The difference in helix angles

of the two wings is about 7◦ , which causes the wing of one rotor to move the material

off the wing of the other rotor, as they approach each other. For uneven rotor speeds, the

44
(a) (b)

(c)

Figure 2.2: Orientation of the rotors for phase angle 45◦ (a), 90◦ (b) and 180◦ (c).

45
interaction between the opposing wings from the two rotors changes with time. The speed

ratios investigated here are 1, 1.125 and 1.5. The left rotor always rotates at 20 rpm, where

the speed of the right rotor varies with the speed ratios. The rotors’ angular positions at

different time instances of a rotor revolution, represented by a cross-section along a plane

in the middle, are shown in Figure 2.3 for the case of speed ratio 1.125. Here 0◦ indicates

the initial position (green) of the rotors. Since the right rotor rotates faster than the left

rotor, when left rotor rotates 90◦ , the right rotor rotates 101.25◦ (blue). Thus, a differential

phase angle of 11.25◦ , exists between the rotors. Similarly a 180◦ rotation of the left rotor

corresponds to 202.5◦ rotation of right rotor (red) and so on. Figure 2.4 shows the change

in the differential phase angles with revolutions for speed ratios 1, 1.125 and 1.5.

2.1.3 Rotor Design

A comparative study of three non-intermeshing rotor-equipped internal batch mixers has

been conducted here to evaluate the performance of the different rotor designs. All the

geometries used here have been taken from United states patents [1, 6, 26], specifications

of which are listed in Table 2.1. The non-intermeshing rotors studied here have either a

different number or configuration of helical wings and are usually chosen based on the

rubber compounding ingredients used and final products. All these three rotors can be

fitted into the mixing chamber without any further modification. Along with the 2-wing

rotor, mentioned in the previous two sections, the other two rotors considered for the study

are 4-wing A and 4-wing B rotors. Both types of the rotors have two helical long wings and

two helical short wings. In the 4-wing A rotor, one long wing and one short wing originate

46
Figure 2.3: Cross-section of the rotor along a plane in the middle at different time instants

during a rotor revolution, represented by different colors, for a speed ratio of 1.125.

Figure 2.4: Differential phase angles Vs. revolutions of the left (slower rotating) rotor

47
(a) 2-wing (b) 4-wing A

(c) 4-wing B

Figure 2.5: Geometry of the three rotors.

48
from one end of the rotor, and another long wing and another short wing originate from the

opposite end as shown in Figure 2.5b, whereas in the 4-wing B rotor, both the long wings

originate from one end and the short wings originate from the opposite end (See Figure

2.5c).

2.2 Materials

The material properties used in the current study are presented in this section, following

previous works [31, 65, 86]. A multiphase technique is adopted to model highly viscous

rubber and air. 75% of the domain is filled with non-Newtonian rubber, with density, ρ1 =

1100 kg/m3 . The rest of the mixing chamber is filled with air, with a density, ρ2 = 1.18415

kg/m3 and a constant dynamic viscosity, µ2 = 1.85508 × 10−5 Pa-s. Rubber resides at

the bottom 75% of the chamber. 75% is a commonly used fill factor and mimics better a

practical situation in comparison to fully-filled chamber [87]. Figure 2.6 shows the average

velocity of rubber for fully-filled and 75% filled conditions. The velocity of rubber in the

75% filled case is higher, and this is expected due to the larger space available for the

material to move. Besides, due to the high viscosity of the rubber, there is higher amount

of plug flow in the 100% filled chamber. Thus, prediction of the distributive and dispersive

mixing statistics will not be a true representation of a real situation, if the chamber is

assumed to be fully filled. In reality, the rubber compound contains ingredients such as

antioxidants, filler, oil etc, which are critical to the desirable properties in the final product,

and the entire mixing process is carried out in multiple passes. However in the current

study, the rubber compound is assumed to be homogeneous.


49
Figure 2.6: Comparison of average velocity of rubber between 100% and 75% fill factors

for 5 revolutions.

50
Table 2.1: Rotor Geometry Specification

Geometric Specification 2-wing 4-wing A 4-wing B

L/D 1.58 1.58 1.58

1st long wing helix angle 38◦ 32◦ 30◦

2nd long wing helix angle 45◦ 40◦ 30◦

1st long wing twist angle 97◦ 90◦ 70◦

2nd long wing twist angle 109◦ 109◦ 72◦

1st long wing length ratio 0.73 0.79 0.66

2nd long wing length ratio 0.64 0.71 0.67

1st short wing helix angle - 32◦ 48◦

2nd short wing helix angle - 40◦ 48◦

1st short wing twist angle - 37◦ 65◦

2nd short wing twist angle - 45◦ 68◦

1st short wing length ratio - 0.34 0.31

2nd short wing length ratio - 0.3 0.33

Originating angular position of 1st long wing 0◦ 0◦ 0◦

Originating angular position of 2nd long wing 178◦ 180◦ 180◦

Originating angular position of 1st short wing - 135◦ 90◦

Originating angular position of 2nd short wing - 314◦ 270◦

51
CHAPTER III

MATHEMATICAL FORMULATION

In the numerical simulations presented here, the laws of conservation of mass and mo-

mentum are solved using a finite volume technique for incompressible transient flows. The

governing equations including continuity (Equation 3.1) and the momentum (Equation 3.2)

equations are expressed here in continuous integral form:

I
v · da = 0 (3.1)
A
Z I I I Z

ρv dV + ρv ⊗ v · da = − pI · da + Tl · da + (fg + fs ) dV (3.2)
∂t V A A A V

In the above equations, V is the volume, A is the surface area, da is the differential face

area, t is the time, v is the velocity vector, p is the pressure, ρ is the density, Tl is the

viscous stress tensor, fs is the surface tension force, fg is the body force due to gravity, I

is the identity tensor and ⊗ is the convolution operator. The viscous stress tensor Tl for

laminar flow is given by:

Tl = 2µ(γ̇)D (3.3)

1
∇v + ∇vT and γ̇ is the shear
 
Here D is the rate of deformation tensor, defined as, D = 2
p
rate, defined as, γ̇ = (2D : D).

The viscosity of non-Newtonian rubber considered here is shear dependent and

Carreau-Yasuda model was found to be best suited in depicting the rheological behavior

52
of rubber [88–90]. The viscosity of rubber, according to Carreau-Yasuda model, can be

defined as:

µ(γ̇) = µ∞ + (µ0 − µ∞ ) (1 + (λ γ̇)2 )(n−1)/2 (3.4)

Here the power constant n = 0.4, the zero-shear viscosity, µ0 = 100, 000 Pa-s, the infinite-

shear viscosity, µ∞ = 1 Pa-s, and the relaxation time constant λ = 10 s. Figure 3.1 shows

the relation between viscosity and shear rate.

Figure 3.1: Change of non-Newtonian viscosity of rubber with shear rate based on Carreau-

Yasuda model

The VOF method has been employed in the current study in order to model the

free surface existing between rubber and air. In the VOF method, both the phases are

53
defined over the whole domain, and are assumed to be immiscible. The fluid properties are

obtained from functions of properties of its constituent phases and their volume fractions,

such as,

ρ = ∑ ρi αi (3.5)
i

µ = ∑ µi αi (3.6)
i
Vi
αi = , (i = 1, 2) (3.7)
V

where, ρi , µi , and Vi are the density, dynamic viscosity and volume of the ith phase. Here

volume fraction α1 corresponds to the amount of rubber phase, while α2 corresponds to the

air phase so that, α1 + α2 = 1. So control volumes with α1 = 1 are fully filled with rubber

and control volumes with α2 = 1 are fully filled with air. The free surface is constituted of

control volumes with 0 < α1 < 1, which usually takes value of α1 = 0.4 to 0.5 [19].

The solutions of the governing equations are obtained for each phase sharing the

same basic principles of mass and momentum transport. The transport equation for the

volume fraction, αi , where i = 1 and 2, is given by:

d
Z Z
αi dV + αi v · da = 0. (3.8)
dt V S

54
CHAPTER IV

COMPUTATIONAL DETAILS

(a) (b)

(c) (d)

Figure 4.1: 2D unstructured quadrilateral mesh for phase angle 45◦ (a), 90◦ (b) and 180◦

(c). (d) Closer view of mesh at the tips.

2D and 3D, unsteady, incompressible and isothermal numerical simulations are

55
conducted using a CFD code, for partially filled mixing chamber with rubber compounds.

Initially rubber is assumed to occupy the bottom 75% of the chamber and top 25% is

occupied by air. For each case, a finite volume mesh is generated for the fluid domain

surrounding each rotor and also for the general fluid domain outside those domains. For

2D phase angle study, finite volume mesh constitutes of unstructured quadrilateral cells

(See Figures 4.1). To ensure sufficient resolution of the mesh near the wall, four prism

layers are created on the surface of the rotors and chamber. A polyhedral mesh gives much

better mesh resolution with less number of cells. So for big 3D geometries, finite volume

unstructured meshes made up of polyhedral, prism and surface elements (See Figure 4.2 &

4.3) are used to reduce total number of cells, in turn reducing simulation time. Adequate

number of cells are created in the clearance between the rotor tip and the chamber wall, and

in the gap between the two rotor tips at the middle. A mesh-independence study was carried

out for three different grid resolutions to choose the appropriate resolution. The time step

was chosen prudently to resolve the motion resulting from the shortest distance traveled

by the rotor based on its rotational speed. A lower time step increases the simulation time

significantly without considerable improvement of results, while greater time step can lead

to numerical instability and unphysical results.

In order to incorporate the rotational motion for the counter-rotating rotors in a

stationary chamber without remeshing the entire domain at every time step, the moving

mesh method was used. In the moving mesh method, fluid region encompassing the ro-

tors rotates, while rest of the fluid region stays steady, thereby creating a direct interface

between these fluid regions such that mass can transfer through the interface.

56
A no-slip boundary condition was used on the rotors’ surfaces and chamber walls,

such that the velocity of the fluid at the rotors surface is equal to the rotor velocity, and the

velocity of the fluid on the stationary chamber wall is zero. Since the rubber considered

here is highly viscous, the flow Reynolds number, Re, is really low and hence the flow is

assumed to be laminar. So, Re = ρUt Dr /µ < 10 based on rubber properties, where ρ is

the density of the fluid, µ is the dynamic viscosity, Ut is the tangential velocity of the rotor

tip, and Dr is the maximum diameter of the rotor. For the uneven rotor speed cases, the

maximum speed was used for the Re estimation. The simulations also take into account the

effect of surface tension and gravity. Compared to the high viscous force, generated from

high viscosity of rubber, surface tension plays a little role [86].

To solve for the transport equations for pressure and velocity, a segregated flow

solver was used. The SIMPLE algorithm is used to solve for the pressure-velocity coupling.

The spatial and temporal discretization of Equation 3.2 is second order and first order

accurate respectively [91]. An Eulerian multiphase model is used to solve the multiphase

flow of air and rubber, where air and rubber have their own velocity field but share the same

pressure field. To obtain a sharp interface between two immiscible fluids such as air and

rubber, the high-resolution interface capturing (HRIC) scheme is used for convection terms

in the VOF transport equation [92]. In addition, continuum surface force (CSF) model is

used for the calculation of the surface tension force [93].

Furthermore, in order to study the dispersive and distributive mixing capabilities,

massless particles were used. Massless particles are transported by the continuous phase,

and hence their velocity is equal to that of the surrounding fluid, v p = v. A Lagrangian

57
multiphase model is used to track the massless particles. These particles were inserted

in the rubber phase initially, and hence it is ensured that they stay in the rubber phase

throughout the simulation.

58
(a) 2-wing (b) 4-wing A (c) 4-wing B

Figure 4.2: Polyhedral mesh on the rotors.

(a) 2-wing (b) 4-wing A

(c) 4-wing B

Figure 4.3: Mesh on the cross-sections.

59
CHAPTER V

RESULTS

5.1 Mixing Measures

5.1.1 Mixing Index

A commonly-used tool to quantify dispersive mixing is called mixing index (See for in-

stance [19, 31, 32, 36, 79, 94]), defined by Dr. Manas-Zloczower as follows [35]:

|γ̇|
λMZ = (5.1)
|γ̇| + |ω|

where |γ̇| and |ω| are the magnitude of the rate of shear and the vorticity tensors, respec-

tively. The mixing index, λMZ , ranging from 0 to 1, characterizes the elongational and rota-

tional components of a flow, with 0 for pure rotational, 0.5 for simple shear, and 1 for pure

elongational flow [20]. Studies have found that for a highly viscous fluid like the rubber

considered here, elongational flows are more effective in droplet as well as solid agglomer-

ate dispersion than simple shear flow, [33,95]. However, mixing index is not time invariant

and multiple times need to be considered within a revolution for an appropriate analysis.In

fact, mixing index is more like a global quantity that discriminate between various designs

and operational parameters, and does not provide enough information about local defects

in the flow field [79]. To assess dispersion, one has to consider the shear rate along with

the mixing index, since higher shear is also needed for effective dispersion [33, 96].
60
5.1.2 Cluster Distribution Index

One of the methods to quantify the goodness of mixing is to compare the calculated mixture

distribution with the ideal one. A statistical quantity to measure the difference between

the two distributions is known as cluster distribution index. The cluster distribution index

indicates the quality of mixing of a cluster of ingredients in a fluid domain [36]. The cluster

distribution index can be calculated using a discrete pairwise correlation function defined

as [20]:
Z r+∆r/2
2
f (r) = ∑ δ (r0i + r)δ (r0i ) = c(r)dr (5.2)
N(N − 1) i r−∆r/2

where f (r) is the coefficient of the correlation function for distance between particle pairs

and thus represents the probability of finding a neighboring particle at the range of r − ∆r/2

to r + ∆r/2 of the ith particle located at ri0 . δ (r) = 1 indicates that the particle is present,

and otherwise, δ (r) = 0. c(r) is the coefficient of the probability density function of the

correlation function, f (r), and the area under curve of c(r) is constant, irrespective of the

shape of the curve.


r=rmax
∑ c(r)∆r = 1 (5.3)
r=0

Here rmax is the largest dimension of the geometry. So no correlation exists between the

particles separated by distance larger than rmax , i.e. c(r > rmax ) = 0. Hence, the cluster

distribution index, ε, can be defined as:


Z ∞
[c(r) − c(r)ideal ]2 dr
0 Z
ε= ∞ (5.4)
2
[c(r)ideal ] dr
0
where c(r) represents the calculated distribution and c(r)ideal , the ideal distribution. The

value of ε depends on the initial position of the particles and the number of the particles.
61
With all the particles clustered together initially (at time t = 0), as is the case in this paper,

ε will have a relatively large value. With time, as the particle distribution gets closer to the

ideal distribution, ε is expected to decrease. Therefore, ideally, ε should tend to 0 [20].

The value of ε is in fact sensitive to the number of the particles and the initial

position of the particles. Sufficient number of particles is necessary to ensure accuracy.

Since all the particles are clustered together initially in the form of a squared, ε will have a

very large value at the beginning of the simulation.

5.1.3 Axial Distribution

The contribution of the axial drag flow created by the helix wing can be estimated using

axial distribution. One of the major components in the distribution process results from

the transfer of material in the middle to the front and back of the chamber, or in other

words, the axial distribution. Axial distribution represents the homogeneity of the particles

distributed on the planes perpendicular to the axial direction or axis of the rotation of the

rotors (i.e. the z-axis) [97]. In order to calculate axial distribution, pairwise correlation

functions have been obtained along the z-axis.

The probability density function of pairwise correlation function was then com-

pared with the ideal distribution. The equations for calculating the axial distribution index

are similar to the cluster distribution index, and are given by:

Z z+∆z/2
2
f (z) = δ (z0i + z)δ (z0i ) = c(z)dz (5.5)
N(N − 1) ∑
i z−∆z/2

62
and Z ∞
[c(z) − c(z)ideal ]2 dz
0 Z
εz = ∞ (5.6)
[c(z)ideal ]2 dz
0

5.1.4 Scale of Segregation

To quantitatively analyze mixing characteristics, Danckwerts defined a quantity named

scale of segregation, which closely relates to the properties of mixture [98,99]. Tadmor and

Gogos used the scale of segregation to describe the texture of the mixture [18]. Connelly

and Kokini also used this parameter to quantify the distributive aspect of mixing [36]. The

scale of segregation for a binary mixture of fluids can be defined as [18]:

Z ζ
Ls = R(r)dr (5.7)
0

where
0 00
∑N
i=1 (xi − x) (xi − x)
R(r) = (5.8)
NS2

and
M
2
∑ j=1 x j − x
S2 = . (5.9)
M

R(r) is called the coefficient of correlation and measures the degree of correlation between

the concentration of two particles separated by distance r. A value of R(0) = 1 implies

that the two particles have the same correlation separated by zero distance, while R(ζ ) = 0

means that there is no correlation between the particles separated by distance ζ . xi0 and xi00

represent the concentration of the ith pair, where x̄ indicates the average concentration. M

is the total number of particles and N is the total number of pairs. S2 is the variance and

calculated from the concentration of all the particles. j indicates index of each particle. Ls

63
is a global measure of the segregated region and does not identify the local defects of the

flow. It is limited by the resolution, which represents smallest length scale within which

concentration difference can be distinguished [18, 36].

5.1.5 Length of stretch

Another parameter that is frequently used to quantify distributive mixing and distinguish

between various mixer designs is the length of stretch [35]. Length of stretch can be defined

as the ratio of distance between a pair of particles at time t, to the distance of same particles

at initial condition, i.e. at t = 0. The length of stretch l can be written as:

|x|
l= (5.10)
|x0 |

where |x| is the distance between two particles at time t, and |x0 | is their initial distance

[100]. Effective mixers shows continuous increase of the length of stretch throughout the

fluid region with time [79].

64
5.2 Phase Angles

5.2.1 Flow Analysis

Figure 5.1 shows the pressure contours for the different phase angles at two different in-

stants of a revolution after 19 revolutions of the rotors. Maximum pressure occurs always

at the front of the rotor tip, while minimum pressure occurs at the back of the rotor tip.

Pressure gradient at the rotor tip depends on the relative position of the tip with respect

to the chamber wall. The pressure gradient is maximum when the rotor tips are inside the

left chamber lobe or the right chamber lobe. Due to higher pressure gradient, dispersion of

the fillers occur at the rotor tip and around the tip corners. It is also noticeable that higher

pressure, which will favor dispersion, occurs at 84◦ rotation as compared to 360◦ . Among

the different phase angles, the materials in the case of 180◦ experience the highest pressure

as the rotor tips come together at the middle, which again is more favorable for dispersion.

The flow behavior in the entire domain, which mainly depends on the rotors’ ge-

ometry and their relative orientations, can be analyzed with the help of velocity contours.

Velocity contour lines for three phase angles are shown in Figure 5.2 at 84◦ and 360◦ ro-

tations of the rotors after 19 revolutions. The velocity contour lines are colored based on

the velocity magnitude. Contours show that the materials are moving in a circular motion

following the rotation of the rotors. The velocity is maximum along circular line created

by the rotor tip. When the rotor tips are at the center of the two chamber lobes, material

exchange occurs. This is evident from the velocity contour lines around the rotor tip when

it wipes of the material from the other rotor causing the material to transfer to the other

65
(a) (b) (c)

(d) (e) (f)

Figure 5.1: Pressure contours for phase angles 45◦ (left), 90◦ (middle) and 180◦ (right) at

84◦ (top) and 360◦ (bottom) rotations of the rotors after 19 revolutions.

66
chamber lobe [See figure 5.2a, 5.2b, 5.2e].

The exchange of materials between the two chambers helps in distribution. The

flow is more streamlined in the 180◦ phase angle case, and thus plug flow is dominant

here. For the 45◦ and 90◦ phase angle cases, the circular streamline pattern is interrupted

by the rotor tip and hence the flow is less streamlined in these cases compared to the 180◦

phase angle case. At 84◦ position of the rotors, for the case of 180◦ phase angle, the two

rotor tips come together at the center and squeeze the materials, thus potentially helping in

dispersion.

(a) (b) (c)

(d) (e) (f)

Figure 5.2: Velocity contours for phase angles 45◦ (left), 90◦ (middle) and 180◦ (right) at

84◦ (top) and 360◦ (bottom) rotations of the rotors after 19 revolutions.

67
5.2.2 Dispersive mixing

A proper dispersion of fillers and other ingredients, through break up of the liquid droplets

and solid agglomerates, is essential in order to obtain homogeneous mixing. This gradual

reduction of filler and agglomerate size by overcoming the cohesive forces, is a critical step

in any mixing process. Many properties of the final product such as tear, abrasion and wear

resistances, and tensile strength depend on the degree of dispersion [101]. A quantitative

study of the filler dispersion is of great importance due to the complexity and difficulty

in dispersing fine and small structured particles of fillers like carbon black and silica [31].

Even though the methods available for quantifying dispersion, have their limitations, the

quantities such as mixing index and maximum shear stress described here, together can

provide valuable information regarding mixing mechanism.

5.2.2.1 Mixing Index

Figure 5.3 shows a histogram of mixing index that indicates the percentage of cells con-

taining only rubber for ranges of mixing index. The histogram was calculated over a period

of 15 rotor revolutions of the rotors. There is a high percentage plug flow in all the phase

angle cases, indicated by low mixing index values (λMZ < 0.1). This is mainly due to the

2D simulation. Since there is no axial movement of the fluid, a large amount sticks to the

rotor wall and rotates with it. There is still a high number of cells with λMZ values that

fall between 0.1 to 0.4, which indicates less dispersive mixing capability. This is especially

evident in the case of the 45◦ phase angle orientation with approximately 39.5%. Within

this range the flow is mainly a combination of plug flow and shear flow, with plug flow

68
being dominant. The fluid near the surface of the rotor primarily experiences this range of

mixing index values, resulting in poor dispersive mixing. For a more effective dispersive

mixing behavior, these fluid elements need to move to the high shear stress regions. The

180◦ phase angle case has the least number of cells (approximately 47%) in the range of

0 < λMZ < 0.3 followed by 90◦ (approximately 48%). 45◦ phase angle, on the other hand,

has highest number of cells in that range (approximately 50%). Thus, materials in 180◦

phase angle has least possibility to experience plug dominant flow and 45◦ phase angle has

highest possibility.

In addition, a substantial percentage of the mixing index values exists between

0.4 and 0.7 for all the three cases, which in turn characterize a shear flow behavior. More

specifically, the percentage values are 33.5, 32.5 and 32 for the 180◦ , 90◦ , and 45◦ phase

angles, respectively.

Furthermore, in the mixing index range of 0.7 to 1, elongational flow is dominant,

which is the most effective for dispersion [33,96]. Elongational flow is mostly experienced

by fluid elements at the gap between the rotor tip and chamber wall, and also in the middle

of the two rotors, when both the rotor tips squeeze materials. The percentage of cells

experiencing elongational flow is very low with the 90◦ phase angle case having the highest

percentage (≈ 9.5%), closely followed by 180◦ case (≈ 8%) and the 45◦ case, which has

the least value of 7.5%. However for effective dispersion, these elongation forces have to

be accompanied by strong shear rates. Therefore, to conclude the observations from Figure

5.3 related to the mixing index calculations, the phase angle of 180◦ and 90◦ perform better

than 45◦ , with 90◦ emerging as the winner by a slight margin. The 180◦ case has the highest

69
percent of cells in the mixing index range 0.4 to 0.7, simple shear, and also high percent

of cells (slightly less than that of the 90◦ case) in the elongational flow range. The 45◦

case performs the worst with the highest percentage of plug flow and lowest percent of

elongational flow.

Figure 5.3: Histogram of mixing index calculated over last 15 revolutions for phase angles

45◦ , 90◦ and 180◦ .

5.2.2.2 Maximum shear stress

Along with elongation, primarily quantified by mixing index, shear stress evaluation is

essential as a measurement of the dispersion effectiveness in the domain [33, 96]. Shear
70
stress is generally higher in the middle of the two rotors and the gap between rotor tip and

chamber wall [33]. Reduction of fillers and other ingredient agglomerates is a function

the shear stress they experience. So the higher is the stress experienced, the greater would

be the amount of dispersed fillers. Thus, to fully understand the dispersive mechanism

of a design, it is imperative to evaluate both mixing index and shear stress magnitude. A

cluster of massless particles is injected into the domain for this kind of evaluation. These

particles are tracked by a Lagrangian multiphase model, and are transported by the con-

tinuous phase, and hence their velocity is equal to that of the surrounding fluid, v p = v.

These particles were inserted in the rubber phase initially, and hence it is ensured that they

stay in the rubber phase throughout the simulation. Figure 5.4 represents the cumulative

probability distribution function of maximum shear stress experienced by the particles for

15 revolutions. The further the curve is towards the right, the higher is the possibility for

the particles to experience larger maximum shear stress, and thus better is the dispersion.

So from the figure it is clear that the 90◦ phase angle case is the furthest towards the left,

and 180◦ is furthest towards the right, implying that the 180 ◦ phase angle case performs

the best in terms of dispersive mixing. In addition, if the critical shear stress required for

breaking a certain kind of filler is knows, it is possible to predict what percentage of that

filler would be broken down at the end of mixing [97]. For example, if 100 kPa is the

critical shear stress, then in the 180◦ phase angle case, at least 30% of the fillers will be

broken down, and 22% and 10% would be the corresponding numbers for the 45◦ and 90◦

phase angles, respectively.

71
Figure 5.4: Cumulative probability distribution function of maximum shear stress experi-

enced by particles.

72
5.2.3 Distributive mixing

The distributive mixing of a cluster of particles is dependent on their initial position. In

order to obtain an unbiased assessment of distributive mixing for the three phase angles,

clusters of particles were injected at three different positions. Each cluster is of the same

size and contains 2500 particles each. The three clusters were tracked through out the

simulation and statistics such cluster distribution index and interchamber material transfer

were calculated.

(a) (b)

Figure 5.5: (a) Initial position of the particle-clusters (magenta circular-middle, cyan

diamond-rotor tip gap, black triangle-left center near rotor wall), (b) Particle-cluster di-

mension.

5.2.3.1 Particle Distribution

As mentioned above, three clusters of massless particles are injected at the beginning of the

simulation at three different locations. Figure 5.5 shows the initial positions of the three

73
particle clusters and also dimension of the cluster. Figure 5.6 shows the particle distribution

of the three clusters for the three phase angles after 5, 15 and 20 revolutions. In general,

most of the particles are homogeneously distributed over time, while some particles do

stick to the chamber wall. The cluster-1, which is initially in the middle of the chamber,

is distributed into the two chambers after 5 revolutions, with the 90◦ phase angle case

having more clustered particles near the left rotor wall. This is probably due to this cluster

being pushed to plug flow zone by the right rotor. After 20 revolutions, the cluster is more

uniformly distributed in the 45◦ and 180◦ phase angles compared to the 90◦ phase angle.

Cluster-2 was placed initially at the gap between rotor tip and chamber wall, and

due to the high shear and elongation experienced in this region, these particles were dis-

tributed more efficiently and evenly with time. Although after 5 revolutions 180◦ phase

angle had most of the particles in the left chamber, there are less differences between the

three phase angle cases after 20 revolutions. Cluster-3 on the other hand, being placed

initially at the plug flow zone, was very poorly distributed for all the three cases, with the

90◦ phase angle case performing the best. Due to the interference of the right rotor tip with

streamlines of the left chamber, cluster-3 was better distributed in the 90◦ phase angle case.

The phase angle of 45◦ performs the worst, as cluster-3 always followed the streamlines

near left rotor, without transferring any particles to the right chamber.

So in summary, from the particle distribution snapshots after 20 revolutions, the

phase angles of 90◦ and 180◦ seem to have more uniformly distributed particles compared

to that of the 45◦ phase angle for all the three cluster positions. Further analysis of dis-

tributive mixing is presented in the next section, with a quantitative analysis of the cluster

74
distribution.

(a) (b) (c)

(d) (e) (f)

(g) (h) (i)

Figure 5.6: Particle distribution for phase angles 45◦ (left), 90◦ (middle) and 180◦ (right)

after 5, 15 and 20 revolutions (from top to bottom).

5.2.3.2 Cluster Distribution Index

One of the main purposes of any mixer is to uniformly distribute the ingredients throughout

the domain; in other words obtain what is known as an optimum distribution. A frequently

75
used quantity to measure how far the obtained distribution is from the optimal distribution

is the cluster distribution index. Wong, Yang and Manas-Zloczower introduced the concept

of cluster distribution index to evaluate and compare distributive mixing performances of

various designs in rubber mixing [33, 37]. Also, Connely and Kokini used cluster distri-

bution index as a measurement of goodness of mixing for clustered ingredients in a high

viscous fluid domain [36, 79]. Cluster distribution index is a statistical quantity that mea-

sures the difference between ideal/optimum distribution and obtained distribution. It is

possible to estimate ideal distribution for any design by assuming all the particles in the

cluster are uniformly distributed throughout the domain. Note that the ideal distribution is

unique, and for the most part, is not time invariant. The difference in the ideal distribution

at different instants of a revolution is so small that, it can actually be neglected. Further-

more, in the current study for the three phase angles, even though the ideal distribution

does not vary more than 1%, obtained distribution of each phase angle case was compared

with the respective ideal distribution for the calculation of cluster distribution index.

In the current study, cluster distribution index has been calculated and plotted

with time for all the three clusters, as shown in Figure 5.7. Due to a disproportionately

large value of ε at the initial time, the first revolution is not considered in the figure. As

mentioned earlier, the cluster distribution index converges to a lower value with revolutions

until it reaches an almost constant value. Among the different cases, the one that reaches

the lowest value in the fastest time has the best distributive mixing characteristics. The

oscillations in the curves are due to the circular motion of the particles as they follow the

streamlines.

76
As mentioned earlier, depending on the position of the cluster, value of ε can be

very high or low. For cluster-1, 90◦ phase angle has a very high value at the beginning and

reduces to 0.175 at the end of 20 revolutions. Phase angle 180◦ performs the best reaching

a lowest value of 0.017 at the end of 20 revolutions, while the 45◦ phase angle case falls

in the middle. Cluster-2 on the other hand, being positioned in the gap between rotor tip

and chamber wall, has a very low value of ε for all the three cases. Although, the plots

fall on top of each other for the three phase angles initially, 90◦ phase angle levels out at a

higher value, but the phase angles of 45◦ and 180◦ keep decreasing even at the end of 20

revolutions.

With regard to cluster-3, 45◦ phase angle performs very poorly with a large value

of ε = 20 at the end of the 10 revolutions. This is due the lack of any disruption of the

streamlines by the rotor tip movement. This causes the cluster to follow the plug flow

along the rotor wall. Even after 20 revolutions, the value is as high as 4.84. With help of

an interfering rotor tip, 90◦ phase angle performs the best with a value of 0.088 followed

by 180◦ , with a value 0.63 at the end of 20 revolutions.

Cluster distribution of all the three clusters were calculated taking positions of

all the particles in each cluster at every instant, and then were plotted against revolutions.

Overall cluster distribution index also indicates the mixing between particles from different

clusters. Based on the cluster distribution index for all the three cluster positions, 180◦

performs very well for all the three cluster positions.

45◦ phase angle performs the worst for the cluster-3 position and 90◦ phase angle

performs the worst for cluster-1 position. This can be further confirmed from Figure 5.8,

77
which depicts the cluster distribution index for the three clusters put together. Even though

ε is lower in the beginning for the 90◦ phase angle, ε becomes even lower in 180◦ case

after 13 revolutions. Note that these conclusions are based on 2D simulations and a better

understanding can be obtained from corresponding 3D simulations.

5.2.3.3 Interchamber material exchange

A continuous transfer of materials from one chamber to the other makes the filler and other

ingredients distribution homogeneous. To judge the interchamber mixing capability of the

three phase angle cases, particles (originally from cluster-2) in the left and right chambers

were identified after 10 revolutions. As was noted earlier in the cluster distribution analysis,

particles of cluster-2 became quite homogeneously distributed after 10 revolutions for all

the three cases, since the cluster itself was placed in the gap between rotor tip and chamber

wall. For cluster-2, the cluster distribution index values similar to each other for all the

phase angles considered here. The percentage of particles in one chamber (left/right) that

transfers to the other is calculated every one second for 10 revolutions and plotted in Figure

5.9. Figure 5.9a illustrates materials transferred from left to right chamber. The 180◦

phase angle transfers the highest percentage of materials, ≈ 13%, and the 90◦ phase angle

transfers least percentage of materials, ≈ 6.5%. Figure 5.9b on the other hand shows a

different trend, where the 45◦ phase angle having the highest transfer of 29%, and the 90◦

phase angle has a higher transfer percentage compared to 180◦ . Also note that, even though

the 90◦ phase angle transfers a high percentage of materials from right to left chamber, the

same is not true when transferring materials from left to right. The 180◦ phase angle

78
(a) (b)

(c)

Figure 5.7: Evolution of cluster distribution index for the three phase angles and three

different initial cluster positions: (a) cluster-1 (b) cluster-2, and (c) cluster-3.

79
Figure 5.8: Evolution of cluster distribution index for all the three clusters combined.

80
case however, has almost an equal transfer of materials from one chamber to the other.

Thus, based on these plots, it can be deduced that the 180◦ phase angle exhibits a better

distributive mixing behavior compared phase angle 90◦ and 45◦ .

(a) (b)

Figure 5.9: Transfer of particles from the (a) left to right chambers, and (b) right to left

chambers for three phase angles: 45◦ (green), 90◦ (green), and 180◦ (red).

81
5.3 Speed Ratios

This section presents results from 3D numerical simulations of rubber compound mixing in

a partially-filled chamber equipped with two counter-rotating 2-wing rotors, for three dif-

ferent speed ratios: 1, 1.125 and 1.5. A speed ratio of 1 indicates that the two rotors are at

even speed, whereas the other ratios indicate uneven speeds. A comprehensive analysis is

presented in terms of a quantitative assessment of dispersive and distributive mixing char-

acteristics. Approximately 4300 particles were injected at the beginning of the simulation

in order to calculate dispersive and distributive quantities.

5.3.1 Dispersive mixing

In an effective mixing process, the size of the ingredient agglomerates is continuously re-

duced by breakup of the solid clumps and immiscible droplets, by overcoming the cohesive

bonds that exist. But fine and small structured particles of fillers like carbon black and silica

are difficult to disperse [31], and dispersion of the fillers in the polymer governs properties

such as tensile strength, abrasion, tear and wear resistances [101]. Therefore, a quantitative

study of the filler dispersion is vital for characterizing dispersive mixing.

5.3.1.1 Mixing Index

A joint probability density function (jPDF) of mixing index and magnitude of shear rate

is shown in Figure 5.10. This quantity indicates the percentage of cells containing only

rubber for ranges of both shear rate and mixing index over a period of 8 rotor revolutions

of the left rotor, after 12 revolutions. Eight revolutions are chosen here because, only for

82
8 revolutions of the left rotor, does the right rotor have an integer number of revolutions in

all the speed ratio cases of 1, 1.125 and 1.5. From the figures, it is clear that, irrespective

of the shear rate, a higher density of the mixing index values exists between 0.5 and 0.6 for

all the three cases, which indicates that the flow is dominated by the shear flow. However,

in the 1.5 speed ratio case, these high-density mixing index regions exist for very low shear

rates (< 10 s−1 ). In the cases of speed ratios of 1 and 1.125, shear rate value ranges from

0 s−1 to 60 s−1 . In all these cases, for mixing index values ranging between 0.5 and 0.6

and shear rates ranging between 10 and 60 s−1 , the 1.125 speed ratio case constitutes the

highest percentage of cells (≈ 18%), followed by the even speed case (speed ratio = 1.0)

at 8%. Note that elongational flow (0.6 < λMZ < 1) is known to be most effective for

dispersion [33].

In the results presented here, elongational flow is more dominant for lower values

of shear rates. Specifically, the speed ratio of 1.5 has the highest percentage of pure elonga-

tional flow (≈ 4%), followed by the even speed case (≈ 3%) and then the 1.125 speed ratio

case (≈ 2%). Elongation generally occurs in the middle of the two rotors, when both the

rotor tips squeeze the agglomerate, thus helping in dispersion. However due to low shear

rates, the elongational forces may not be strong enough for effective dispersion.

In addition, there is a high percentage (≈ 44%) of low mixing index values (0 <

λMZ < 0.2) at low shear rates (< 10 s−1 ) characterizing plug flow especially in the 1.5

speed ratio case. The plug flow contribution is observed to be the least for the 1.125 speed

ratio case, which is about 16%. The low mixing index values occur near the surface of

the rotor, as the material rotates along with it, resulting in poor dispersive mixing. For

83
dispersive mixing to be more effective, these fluid elements have to undergo high shear

rates.

So in summary, the results demonstrate that the speed ratio 1.125 performs the

best, since this case gives the lowest percentage of plug flow and the highest percentage

of high shear rates (> 10s−1 ) high mixing index values. Even though, the speed ratio 1.5

has the highest elongational flow, the shear rate is low, and the percentage of high shear

rate at high mixing index values is the lowest. Furthermore, the speed ratio of 1.5 gives the

highest plug flow, thus performing the worst.

5.3.1.2 Maximum shear stress

Even though mixing index provides some valuable insight with regard to dispersive mixing,

it does not indicate the flow strength for effective agglomerate breakup [97, 102]. Besides

mixing index, another important parameter that illustrates the dispersive mixing capability

is the maximum shear stress [33, 39], especially in a shear-driven flow such as the one

presented in this paper, where the rubber viscosity is of the order of 105 Pa-s. The higher the

shear stress experienced by the ingredient, the greater is the possibility for dispersion. The

cumulative distribution plot in Figure 5.11 depicts the maximum shear stress experienced

by the particles over one revolution after 20 revolutions for the three speed ratios studied

here. The further the curve is towards the right, the higher is the probability of maximum

shear stress. Ideally, one would want to consider a period of 8 revolutions for the analysis

of the maximum shear stress for reasons mentioned in the jPDF analysis above. However,

due to the unavailability of data, only results over a period of one revolution are presented

84
(a) (b)

(c)

Figure 5.10: Joint probability density function (jPDF) of the mixing index and shear rate

calculated over a period of 8 rotor revolutions of the left rotor after 12 revolutions for speed

ratios of (a) 1, (b) 1.125 and (c) 1.5.

85
here, and it is assumed that ranking of the three cases does not change whether we look at

one revolution or 8 revolutions. If the critical value required for rupture of particular filler

is known, it is possible to determine the number of particles broken up under shear stress at

least one time [97]. It is apparent from the plot that in the case of speed ratio of 1.5, most of

the particles experienced significantly lower shear stress in comparison to the speed ratio

of 1.125 and the even speed case. For instance, about 33% of the particles experienced

a shear stress higher than 60 kPa for a speed ratio of 1.125, followed by the even speed,

which was about 27%. For the speed ratio of 1.5, less than 5% particles experienced shear

stress higher than 60 kPa.

Figure 5.11: Cumulative distribution of maximum shear stress experienced by the particles

over a period of one revolution

86
5.3.2 Distributive mixing

A good spatial distribution of the broken agglomerates represents good distributive mixing.

The dynamics of distributive mixing is assessed in this study through the investigation of

parameters such as cluster distribution index and scale of segregation. These quantities

can be numerically calculated by tracking the motion of a set of massless particles in the

domain. In addition, the axial distribution of these particles is also investigated in order

to analyze the effectiveness of this feature in assisting towards better distributive mixing

characteristics.

5.3.2.1 Cluster Distribution Index

In the current study, particles are injected in the form of a spherically-shaped cluster for

all the three speed ratios at the beginning of the simulation. An ideal distribution has been

used as a reference for an unbiased comparison of the three cases. The ideal distribution is

obtained by assuming that all the particles are uniformly distributed throughout the domain

and hence it is the most optimum distribution that can be obtained by the mixer. Fig-

ure 5.12a shows the initial spherical cluster distribution of particles relative to the rotors.

The corresponding probability distribution function for the pairwise correlation function

is shown in Figure 5.12b. The values are zero at the ends, which represent minimum and

maximum dimensions of the geometry. The curve has a very sharp peak because, initially

all particles are lumped together in the shape of a sphere, this representing a small particle

pair distance. Also shown on this plot is the probability distribution function of the ideal

distribution. The purpose of any mixer is to spread out the particles and in the ideal case,

87
(a)

(b)

Figure 5.12: (a) Initial position of the particles (b) Probability distribution of the particle

pair distances at the initial condition (red line); Also shown here is the probability distribu-

tion of the particle pair distances of the ideal distribution (black line with symbols).

88
the curve is more flat due to the particles being homogeneously distributed. It should be

noted that the area under both these curves is the same. Also, the differences in the ideal

distribution for a fixed geometry at different times were negligible, and the changes in the

geometry of the three cases resulted in differences < 1% in the ideal distribution. So a

fixed ideal distribution was used for all the three cases for the evaluation of ε.

Figure 5.13 shows the instantaneous particle distribution at three different times,

i.e. after 5, 15 and 20 rotor revolutions, for the three cases studied here. Qualitatively, after

5 revolutions (See Figures 5.13a, 5.13b and 5.13c), the speed ratios of 1.125 and 1.5 have

their particles more spread out in comparison with the even speed case. This is expected

due to the asymmetric nature of the flow field resulting from the uneven speeds, which

do cause some particles to move faster than others. However after 15 revolutions (See

Figures 5.13d, 5.13e and 5.13f) and 20 revolutions (See Figures 5.13g, 5.13h and 5.13i),

the differences between the cases are not as obvious. This indicates that the uneven speeds

do result in a more homogeneous distribution occurring faster than the even speed case. But

with time the distributions between the three cases tend to be similar to each other. These

observations can be further affirmed by quantifying the differences in particle distributions

between the three speed ratios.

Figure 5.14 shows the temporal evolution of the cluster distribution index for the

three speed ratios. Since cluster distribution index depends on the position of the cluster,

the continuous flow field change at the middle of the rotors due to the speed ratio, affects

the mixing performance. The oscillations are due to the circulation of particles in the

flow domain by the stretching and unstretching of the cluster of particles. The oscillation

89
amplitude is lower in the speed ratios of 1.125 and 1.5, since the differential phase angle

of the two rotors interrupts the mostly regular and circular pattern, except very near the

rotor walls, thereby introducing more randomization in the mixing. As mentioned earlier,

the cluster distribution index, ε, is large initially and keeps reducing with revolutions. The

plot does not include the first revolution due to the disproportionately high values resulting

from the particles being lumped together. In first few revolutions, ε reduces faster for both

the uneven speeds (speed ratio of 1.125 and 1.5) compared to the even speed. However, ε

for the speed ratio of 1.5 levels out at a higher value compared to the other two cases at the

end of 20 revolutions. In addition, the speed ratio of 1 and 1.125 decrease with time, with

1.125 converging to the lowest value in the fastest time. Based on the cluster distribution

index, a speed ratio of 1.125 exhibits the best distributive mixing characteristics among all

the three cases.

5.3.2.2 Axial Distribution

Figure 5.15 shows the time evolution of the axial distribution index after the 1st revolution,

again due to disproportional large values at the initial time. The lower the value is for the

axial distribution index, εz , the more uniform are the particles along the axis of rotation.

Clearly, at the start, εz for the even speed case is higher compared to the speed ratios of

1.125 and 1.5. This is because initially all the particles are at the middle and velocity

gradient is higher for speed ratios 1.125 and 1.5, since one rotor rotates faster than the

other. At the beginning of mixing, εz in the speed ratio of 1.5 decreases faster than the

others. However, it reaches a plateau at about 0.25, after 11 revolutions. The even speed

90
(a) (b) (c)

(d) (e) (f)

(g) (h) (i)

Figure 5.13: Particle distribution for speed ratios 1 (left), 1.125 (middle) and 1.5 (right)

after 5, 15 and 20 revolutions (from top to bottom).

91
Figure 5.14: Evolution of the cluster distribution index with time.

case converges at a slightly lower value of 0.24, while the speed ratio of 1.125 continues

to decrease and eventually converges to the lowest value of about 0.20 at the end of the

20 revolutions. This indicates that the speed ratio of 1.125 provides best axial distribution,

and 1.5 the worst.

5.3.2.3 Interchamber material exchange

To obtain better mixing it is also necessary for the ingredients to transfer from one chamber

to the other as the rotors rotate. To evaluate the amount of transfer, particles injected in the

domain initially as a sphere-shaped cluster are tracked. At every ∆t = 1 s, the current

position of a particle is compared with its previous position. If an exchange of a particle

happens between the left and right chambers within ∆t, the transfer is counted.

Table 5.1 presents the total counts as a percentage of total particles transferred at

92
Figure 5.15: Evolution of axial distribution with time.

least once from one chamber to the other, crossing the middle region between the rotors.

From the table, we observe that the percentage of transfer increases with time for all the

speed ratios. However the rate of this increase is the highest with the speed ratio of 1.125

at 58.95% after 20 revolutions, while it is the least for the speed ratio of 1.5 at 42.8%

after 20 revolutions. These results confirm what was observed with the cluster distribution

index, which also showed that 1.125 and 1.5 had the best and worst distributive mixing

characteristics, respectively.

5.3.2.4 Scale of Segregation

In order to illustrate qualitatively the importance of scale of segregation, particles dis-

tributed after 10 revolutions are assigned two different concentrations, values of which are

chosen arbitrarily. The concentrations of particles in the upper half of the chamber are

93
Figure 5.16: Two types of particles with concentration 1 (green) and 0 (black) defined at

the end of 10 revolutions for scale of segregation calculation

(a) (b) (c)

(d) (e) (f)

Figure 5.17: Front view of the distribution of particles for speed ratios 1 (left), 1.125

(middle) and 1.5 (right) at 12 (top) and 20 (bottom) revolutions; The green and black

particles represent concentrations of 1 and 0, respectively.

94
Table 5.1: Percentage of Particles transferred at least once from one chamber to the other

0th revs. 5th revs. 10th revs. 15th revs. 20th revs.

1.0 0% 28.4% 40.4% 43.4% 47.0%

1.125 0% 34.94% 46.5% 53.2% 58.95%

1.5 0% 36.1% 41.7% 42.35% 42.8%

assigned as 0, whereas those in the lower half are assigned 1, represented by the black and

green particles, respectively, as shown in Figure 5.16. From the front view of the rotors

in Figures 5.17a, 5.17b and 5.17c, it is clear that two revolutions after the particles were

assigned dual concentrations, most of particles are still segregated for speed ratios of 1 and

1.5, with 1.5 showing more. As time progresses, more integration between the particles

does occur for all the cases as seen after 20 revolutions in Figures 5.17d, 5.17e and 5.17f.

However, the particles in the speed ratio of 1.5 are still quite segregated, and more segre-

gated than both 1.125 and 1. Furthermore, Figure 5.18 shows the green and black particle

distribution in a top view for all the cases at two times: 12 and 20 revolutions. Here it

can be seen that particles in the middle are mostly mixed as a result of the interchamber

particle exchange (See Figures 5.18a, 5.18b and 5.18c), especially for speed ratios of 1.125

and 1.5. This is because the uneven speeds create a velocity gradient in the middle. Later

after 20 revolutions, there is more randomness and less segregated regions in the particles

distributed, as the materials are getting stretched and folded (See Figures 5.18d, 5.18e, and

95
5.18f). Overall, the speed ratio of 1.125 has the least segregated regions, especially in the

middle and near the rotor surface (See Figures 5.17b and 5.18b), and the speed ratio of 1.5

has the most segregated regions, thus failing to transfer material in the upper and lower

halves. It has less segregated regions at the middle due to the velocity gradient, but most of

the particles at the top and bottom are segregated (See Figures 5.17f and 5.18f). From these

figures it is expected that the speed ratio of 1.5 will have the highest segregation scale and

1.125 will have the least segregation scale at the end of the simulation. One of the reasons

that the 1.5 speed ratio case exhibits such a behavior is that one rotor rotates much faster

than the other, and the the chamber with the slower rotor is unable to take in material, since

it cannot push out the materials fast enough.

Figure 5.19 shows the evolution of the scale of segregation with revolutions. The

initial value of scale of segregation is higher due to the initial setup of particles (Figure

5.16) in the upper and lower halves. As the particles get mixed with time, scale of segre-

gation reduces and converges to a lower value. After 1 revolution (at 11 revolutions on the

x-axis), the 1.125 speed ratio case has the lowest value and continues to further decrease,

whereas the 1.5 speed ratio has the highest value of scale of segregation. The evolution

of scale of segregation with revolutions also shows oscillations. This can be explained by

the movement of particles in the rotating flow region causing stretching and unstretching

of the material. The speed ratio of 1.125 shows the least oscillations compared to the other

two speed ratios because, among all the three cases, the 1.125 case takes the longest time

(8 revolutions of the left rotor) for the rotor-to-rotor interaction pattern to repeat. After 10

revolutions, the segregation scale of all the three speed ratios converge to an average value,

96
with the speed ratio of 1.125 having the lowest value of 0.007, followed by the speed ratio

of 1 with 0.03, and then the 1.5 speed ratio case with an average segregation scale value

of 0.14. This confirms the qualitative picture of segregated particles seen in Figures 5.17

and 5.18, where the largest regions remained segregated even after the mixing for the 1.5

speed ratio case. On the other hand, the speed ratio of 1.125 exhibited superior distributed

mixing compared to other two speed ratios.

(a) (b) (c)

(d) (e) (f)

Figure 5.18: Top view of the distribution of particles for speed ratios 1 (left), 1.125 (mid-

dle) and 1.5 (right) at 12 (top) and 20 (bottom) revolutions; The green and black particles

represent concentrations of 1 and 0, respectively.

97
Figure 5.19: Evolution of scale of segregation with time.

To conclude the analysis of the dispersive and distributive mixing capabilities of

the three speed ratios, Figure 5.20 presents a bar graph of the major performance param-

eters including, elongational flow components, maximum shear stress, and plug flow for

dispersive mixing, and cluster distribution, axial distribution, inter-chamber particle trans-

fer and segregation scale for distributive mixing. For each parameter, different speed ratios

have been ranked from 1 to 3 based on performances. Since the plug flow contribution

is a negative attribute with regard to mixing characteristics, it is indicated on a negative

scale. With regard to dispersive mixing the 1.125 case is superior due to its least plug flow

contribution and highest maximum shear stress ranking. While the 1.5 case does have the

largest contribution of elongation flow components, as mentioned previously, these were at

low shear-rates. Due to this reason and also the fact that it has the highest plug flow con-

tribution, the 1.5 speed ratio has the worst dispersive mixing characteristics. On the other

98
hand, for distributive mixing, 1.125 is consistently the best in all performance parameters

and 1.5 is the worst.

(a) (b)

Figure 5.20: Summary of (a) dispersive mixing and (b) distributive mixing capabilities for

three speed ratios: 1 (blue), 1.125 (red), and 1.5 (green).

99
5.4 Rotor Design

Numerical simulations of three different types of counter-rotating rotors in a mixing cham-

ber partially-filled with rubber compounds have been conducted. The rotor types are 2-

wing, 4-wing A and the 4-wing B rotor. The mixing chamber is taken to be partially-filled

with 75% rubber and both the rotors have the same speed of 20 RPM. The primary objec-

tive of this study is to assess and compare dispersive and distributive mixing characteristics

between the rotor types via various quantities such as flow patterns, flow rates, and other

mixing quantities such as cluster distribution index and length of stretch.

5.4.1 Flow patterns and flow rates

Favorable material movement enhance effective mixing and result in an improved perfor-

mance. Velocity vectors and flow rates are analyzed in the entire domain through various

cross-sections as shown in Figure 5.21, after blanking out the air. Figures 5.22 and 5.23

show the velocity vectors through cross-sections of Figure 5.21c at two different times,

corresponding to 1.2 and 1.67 revolutions. Figure 5.24 shows the velocity vectors through

the axial cross-sections at two different times, again corresponding to 1.2 and 1.67 revolu-

tions. These times were chosen for clearer visualization of the typical material movement,

which solely depend on the different rotor designs. The velocity vectors are colored by

velocity magnitude and big arrows are employed to demonstrate the directions of resulting

bulk movement. A lot of information about material movement can be further affirmed by

analyzing flow rates across sections. Studies have been previously conducted with flow

rate analyses for understanding the effect of rotor design and fill factor [85]. Flow rate

100
plots corresponding to the 15-17 revolutions are shown here, so that their magnitudes are

unaffected by the initial condition. A change of flow rate from positive to negative and vice

versa within different phases of a revolution indicates the rubber movement direction and

magnitude. Normalized flow rates across cross-sections shown in Figure 5.21 and colored

by the same cross-sections, are presented in Figures 5.25 (colored by Figure 5.21a) and

5.26 (colored by Figure 5.21b).

(a) (b) (c)

Figure 5.21: Cross-sectional planes along the (a) axial, and (b) and (c) transverse directions

for analysis of flow rates (using (a) and (b)) and velocity vectors (using (c)) at time, t=0.

For the 2-wing rotor at 1.2 revolutions and cross section at 25% (See Figure 5.22a),

there is a transverse movement of the material from the right chamber cavity to the left

chamber cavity as the wings give push, and in the opposite direction at 1.67 revolutions

(Figure 5.23a). The flow is vice-versa for the 75 % cross section as shown in Figures 5.22g

and 5.23g. So at the 50% cross-section (in Figures 5.22d and 5.23d) the transverse material

movement in the two directions results in the material being squeezed down. Axial flow
101
(a) (b) (c)

(d) (e) (f)

(g) (h) (i)

Figure 5.22: Instantaneous velocity vectors colored by velocity magnitude for 2-wing (left),

4-wing A (middle) and 4-wing B (right) for cross-sections at 25%, 50% and 75% (from top

to bottom) of the rotor length at 1.2 revolutions; cross-sections shown in Figure 5.21c.

102
(a) (b) (c)

(d) (e) (f)

(g) (h) (i)

Figure 5.23: Instantaneous velocity vectors colored by velocity magnitude for 2-wing (left),

4-wing A (middle) and 4-wing B (right) for cross-sections at 25%, 50% and 75% (from top

to bottom) of the rotor length at 1.67 revolutions; cross-sections shown in Figure 5.21c.

103
patterns in Figure 5.24 show material moving forward and backward at the left and right

sides, respectively at 1.2 revolutions (See Figure 5.24a) and vice versa at 1.67 revolutions

(See Figure 5.24d), thus keeping the material in continuous motion. These flow patterns

are further confirmed from the flow rate evolutions over two revolutions shown in Figures

5.25 and 5.26. For the 2-wing rotor, distinct peaks due to the two long wings are seen in

the flow rate evolutions across cross-sections in the front and the back of the rotor (Figures

5.25a and 5.25g, respectively), whereas in the middle (Figure 5.25d), the transverse flow in

both directions results in negligible net flow. Axial flow patterns across the cross-sections

on the left and right (Figures 5.26a and 5.26g, respectively), show the exact same behavior

as well, and again, in the middle (Figure 5.26d), flow is a lot more random.

Flow patterns of the 4-wing A rotor have some differences compared to the 2-wing

rotor. Due to the additional two small wings 4-wing A rotors has additional fluid movement

which possibly intensifies mixing action in axial and transverse direction (Figures 5.22b,

5.22e, 5.22h, 5.23b, 5.23e, 5.23h,5.24b, and 5.24e). So with regard to the flow rates across

axial cross sections, the peaks from the two long wings are no longer as evident as 2-wing

rotor in the 4-wing rotors case for any of the cross-sections, due to a more simultaneous

transfer of material in both directions (See Figures 5.25b and 5.25h). With regard to axial

flow rates on the other hand, the major flow pattern is very similar to the 2-wing rotor (See

Figures 5.26b and 5.26h).

Distributive mixing in the 4-wing B rotor is mainly achieved by the long wings

in them transferring fluid from one chamber to another (Figures 5.22c, 5.22f, 5.22i, 5.23c,

5.23f,5.23i, 5.24c, and 5.24f), but more interestingly, it can be seen that, because both the

104
long wings originate from the same end in the 4-wing B rotor, material transfers always in

one direction, rather than switching directions at different rotor phases. This can be further

confirmed from the flow rate plots, where the transverse motion of the material across

the axial cross-sections is similar to the 4-wing A rotors for the axial cross-sections (See

Figures 5.25c and 5.25i) and for the transverse cross-sections, the positive-only value at the

left side (blue) and negative-only value at right side (green) of the flow rate plot indicates

that flow almost does not change the direction on the sides of the rotors, again since both

the long wings are on the same side (See Figures 5.26c and 5.26i). Hence, it is possible

that the 4-wing B rotor will have poor dispersive and/or distributive mixing compared to the

other two rotor cases. However, that must be further proven using other more quantifiable

mixing metrics.

5.4.2 Dispersive mixing: Mixing Index & Shear Rate

Dispersion of fillers and agglomerates through a process of breaking them down using

stresses and forces in order to overcome cohesive force and interfacial forces is necessary

for the resulting mixture to be homogeneous. Since elongational flow is more effective

than simple shear, mixing index is a quantity that can determine dispersive performance

of a mixer, specially for a high-viscous fluid [35]. There have been lot of studies, where

mixing index was used as a means to quantify dispersive mixing [19, 31, 32, 36, 79, 94].

However, in order to really assess how effective the shear and elongational forces are,

one has to consider the shear rate in conjunction with the mixing index [33]. Figure 5.27

shows the joint probability density of mixing index and shear rate for rubber alone, using

105
(a) (b) (c)

(d) (e) (f)

Figure 5.24: Instantaneous velocity vectors of 2-wing (left), 4-wing A (middle) and 4-wing

B (right) rotor at 1.2 (top) and 1.67 revolutions (bottom), for the cross section through the

both rotor along their axis.

106
(a) (b) (c)

(d) (e) (f)

(g) (h) (i)

Figure 5.25: Evolution of normalized flow rate of 2-wing (left), 4-wing A (middle) and

4-wing B (right) rotors for 2 revolutions from 15 to 17, through the cross sections shown

in Figure 5.21a; Lines are colored by cross-sections shown in Figure 5.21a.

107
(a) (b) (c)

(d) (e) (f)

(g) (h) (i)

Figure 5.26: Evolution of normalized flow rate of 2-wing (left), 4-wing A (middle) and

4-wing B (right) rotors for 2 revolutions from 15 to 17, through the cross sections shown

in Figure 5.21b; Lines are colored by cross-sections shown in Figure 5.21b.

108
calculated data collected at multiple time instants over a period of one rotor revolution after

17 revolutions.

Mixing index values close to zero and 0.5 indicate plug flow and shear flow, re-

spectively, whereas, values higher than 0.5 and tending towards 1 indicate elongational

flow. From the figures it is clear that the flow is dominated by the shear flow, as a higher

density of the mixing index values exists between 0.5 and 0.6 for varying shear rates for

all the three rotors. Also, the density of elongational flow is very low for any shear rate

for all the rotors, with the 2-wing rotor being marginally better than the other two rotors.

The elongation occurs mostly in the middle of the two rotors when both the rotor tips push

down the rubber, thereby squeezing and breaking the material. Furthermore, there is a rea-

sonably high percentage of low mixing index values (0 < λMZ < 0.2) at low shear rates

(< 10s−1 ) which indicates plug flow and thereby poor dispersive mixing. This kind of be-

havior is exhibited the most by the 4-wing B rotor about 30% and the least by the 4-wing

A rotor about 10%. Typically this is found near the surface of the rotor and along with

which the material rotates without undergoing breakdown of the fillers and agglomerates.

For effective mixing these fluid elements needs to move to a high shear and mixing index

region. So in summary, the feature to be assessed from this figure is the percentage of node

points having the high mixing index and the high shear rate and according to that metric,

the 4-wing A rotor shows the highest percentage of those points followed by the 2-wing

followed by the 4-wing B rotors. The 4-wing A rotor has about 22% of mixing index values

between 0.5 and 0.6 for the highest shear rates ranging from 10-60 s−1 , with the number

for the 2-wing rotor being at 12% for this same shear rate. The 4-wing B rotor has a max-

109
(a) (b)

(c)

Figure 5.27: Joint probability density of the mixing Index and shear rate calculated over a

period of one rotor revolution after 17 revolutions for the (a) 2-wing (b) 4-wing A and (c)

4-wing B rotors.

110
imum shear rate range of 10-40 s−1 with only about 7% of mixing index values between

0.5 and 0.6 experiencing it. The results demonstrate that, based on the metric presented

here, the 4-wing B rotor has the worst, while 4-wing A rotor has the best dispersive mixing

characteristics.

5.4.3 Distributive mixing

In order to study distributive mixing of the three rotors, massless particles were inserted into

the computational domain. These particles are tracked using the Lagrangian Multiphase

model and their positions are calculated at every time step using the Eulerian solution of

the velocity at every grid point. The particles were injected initially within a sphere and

were tracked through a certain time. The data and information obtained from particles are

used to quantify the distributive mixing performance of the three rotors.

5.4.3.1 Particle Distribution

4000 particles were inserted at the middle of the two rotors initially and the snapshots

of the evolution of particle distribution are presented here in Figure 5.28 for a qualitative

evaluation. From the figures, it can be seen that after one revolution of the rotor, some of the

particles (for all 3 rotor types) move to the left, while the rest move to the right due to the

various wing actions. As time progresses and after 10 revolutions, the 2-wing and 4-wing

A rotors show a more homogeneous particle distribution in comparison to 4-wing B rotor.

Again after 17 revolutions it is clear that the 4-wing A rotor shows the most homogeneity

in particle distribution followed by the 2-wing and 4-wing B rotors. This can also be tied

to the highest and lowest shear rate distributions of the shear flow regions in the 4-wing

111
(a) (b) (c)

(d) (e) (f)

(g) (h) (i)

(j) (k) (l)

112

Figure 5.28: Particle distribution for 2-wing (left), 4-wing A (middle) and 4-wing B (right)

at the initial condition, and after 1, 10 and 17 revolutions (from top to bottom).
A and 4-wing B rotors, respectively. The high shear rate for instance is very effective in

breaking down the agglomerate of particles and resulting in the particles being spread out

more in the domain as visible in the 4-wing A particle distribution.

5.4.3.2 Cluster Distribution Index

The particle distribution in the three different rotors are compared to one another using the

ideal distribution of particles as reference [37]. An ideal distribution is obtained based on

the premise that the ultimate goal of any mixing equipment is to spread out all the particles

uniformly throughout the chamber. This distribution solely dependent on the geometry of

the rotor. The ideal distribution is not time invariant but within one revolution the change is

negligible in our case and thus considered time independent. Figure 5.29a shows the initial

distribution of particles, while Figure 5.29b, 5.29c & 5.29d show the simulated distribution

of particles after 17 revolutions along with the ideal distribution for the three rotors. The

minimum particle-particle distance, r, would be zero while the maximum corresponds to

the maximum dimension of the mixing chamber. The two 4-wing rotors have a slightly

different ideal distribution compared to the 2-wing rotor, due to the two additional wings,

but in spite of the short wings originating from different sides, ideal distributions are almost

indistinguishable between the 4-wing A and the 4-wing B rotors. The initial distribution

of particles, which is exactly the same for all the three rotors, has a very sharp peak very

close to a particle-particle distance corresponding to r = 0, since particles were injected

at the very beginning of the simulation in a sphere. As time progresses the particles get

distributed throughout the domain, and the sharp peak of initial condition disappears. On

113
comparing the particle distribution after 17 revolutions with the ideal, the 4-wing B rotor

has largest deviation compared to the 2-wing and the 4-wing A rotors, which have similar

deviations from their respective ideal particle distributions.

This deviation from the ideal distribution can be quantified using the cluster dis-

tribution index parameter defined earlier. Figure 5.30 plots the evolution of the cluster

distribution, ε, with revolutions for all the three rotor designs. It is expected that ε will

usually starts from a high value (depending on the initial distribution) and then decrease

and converge to a value with time. A better rotor design will reach a smaller converged

value in a shorter time. The first revolution has been excluded from the comparisons due

to the disproportionately large value at the beginning of the simulation, resulting from the

particles being packed in a sphere. Plots show that for the 4-wing A rotor, the ε drops to

its minimum (of ≈ 0.07), in the fastest time (≈ 5 revs), whereas, the ε of 4-wing B rotor

decreases to approximately 0.12, after around 10 revs. The ε of the 2-wing rotor decreases

to 0.08 after about 6 revs. These results indicate that the 4-wing A rotor approaches the

ideal distribution in the fastest possible time, thus exhibiting the best distributive mixing

characteristics, while the 4-wing B rotor performs the worst.

5.4.3.3 Length of Stretch

The length of stretch is an indication of the ability of mixers to stretch a fluid element

over time by stretching and folding. The evolution of the mean of the length of stretch for

the three rotors have been plotted with time in Figures 5.31. The mean is calculated from

the particle data corresponding to the material. As the number of particles experiencing a

114
(a) Initial Distribution (b)

(c) (d)

Figure 5.29: Probability distribution of particles (a) at the initial time, and after 17 revolu-

tions for (b) 2-wing, (c) 4-wing A and (d) 4-wing B rotors (solid lines); Also shown are the

corresponding ideal distribution of particles for the three different rotors in (b), (c) and (d)

(lines with symbols).

115
Figure 5.30: Evolution of the cluster distribution index with time.

Figure 5.31: Evolution of the mean length of stretch with time.

116
continuous stretching due to the shear increases with time, the mean of the length of stretch

increases. The oscillations in these figures represent the stretching and unstretching of the

fluid element. The figure shows that the average length of stretch is the highest for the

4-wing A rotor indicating again its best distributive mixing qualities, consistent with the

previous parameters evaluated in this study.

117
5.5 Validation

A mesh independence study was conducted to ensure accuracy and confidence in the results

presented here. Figure 5.32 shows the velocity profile along the x-axis across the center

cross section of the left half rotor of the 2-wing geometry. The mesh independence study

was conducted for 600k, 1.3M and 2M cells at 1.2 revolutions. Assuming the simulation

with 2M cells yields the most accurate results, the error obtained for 1.3M cells case was

about 2%. Hence, all other simulations were carried out using 1.3M cells. In addition, time

step and iteration independence studies were also carried out to assure less numerical error.

Furthermore, the simulation result was qualitatively matched against previously published

Figure 5.32: Velocity profile along x-axis at the mid-cross section of the left half rotor for

600k, 1.3M & 2M Cells.

experimental results. Freakley and Patel [17] performed experiments to investigate flow

and temperature properties in a internal mixer. The Farrel Bridge BR Banbury, which has

118
two wings, was used for this investigation. The rotors were rotated at 50 rpm at the speed

ratio of 1:1.125. 75% of the chamber was filled with rubber. This particular publication

was chosen for experimental validation, since the Farrel Bridge BR Banbury is very similar

to the 2-wing geometry used in the current study. The measured pressure profile from the

study of Freakley and Patel was plotted against the pressure obtained from the simulation of

the 2-wing rotors at 20 rpm even speed [See Figure 5.33]. At each revolution there are two

peaks, as two wings cross over the point of interest (pressure transducer location), twice

within a revolution. The experimental results have higher peak due to the larger dynamic

pressure head resulting from the larger speed. Other discrepancies are due the differences

in the geometry, experimental setup and speed ratio.

Figure 5.33: Comparison of pressure profile obtained from Freakley and Patel’s experiment

at 50 rpm at the speed ratio of 1:1.125 [17] and simulation at 20 rpm even speed.

119
CHAPTER VI

CONCLUSION

A simplified 2D simulation model was developed to analyze the mixing capability of three

phase angles: 45◦ , 90◦ and 180◦ . Later, 3D simulations were conducted for speed ratios

1.0, 1.125 and 1.5. Mixing with three different rotor designs, 2-wing, 4-wing A and 4-wing

B rotors was also investigated to determine their performance. All numerical calculations

were carried out for partially-filled rubber mixing in a rotor-equipped chamber. The cham-

ber was assumed to 75% filled with rubber. The two rotors were rotating at 20 rpm for the

phase angle and rotor design investigations, and the right rotor had higher speed than the

left rotor, in the case of speed ratios of 1.125 and 1.5 for the speed ratio investigations. All

the analysis was performed for evaluating both dispersive and distributive mixing capabil-

ities along with flow patterns.

1. Phase angles: The flow field was analyzed from velocity and pressure maps in the

domain. The velocity contours showed lot of circulation mainly because of the 2D

geometry being considered. The flow was observed to be more streamlined in the

180◦ phase angle case compared to other two. This is because of the absence of any

disruption of the circular streamlines by the rotor tips in the 180◦ phase angle case.

Pressure contours showed that pressure was maximum at the front of the rotor tip

and minimum at the back. Pressure gradient was maximum at the gap between rotor
120
tip and chamber wall. The phase angle of 180◦ had a higher possibility of dispersion,

due to the squeezing of materials, as the two rotor tips meet each other in the middle.

(a) Dispersive mixing of the three phase angle cases was quantified by calculating

the histogram of mixing index of all the cells, and cumulative distribution function

of maximum shear stress experienced by massless particles injected in the domain.

Mixing index showed that elongation was slightly better in 90◦ phase angle compared

to the 180◦ phase angle, and 45◦ phase angle performed the worst, since the latter

case showed both highest plug flow and lowest elongation. Cumulative probability

distribution function of maximum shear stress showed that fillers in the 180◦ phase

angle case had highest possibility to experience a large maximum shear stress, and

90◦ phase angle showed worst possibility of filler breakup. Hence, this analysis

concluded that 180◦ was superior to the others in terms of dispersive mixing.

(b) Distributive mixing was analyzed qualitatively by observing particle distribu-

tion snapshots, and quantitatively by considering cluster distribution index and in-

terchamber material transfer. Since the results very much depended on the initial

position of the cluster, three different positions were chosen for particle-cluster in-

jection in the current study in order to provide a complete and unbiased judgement

of the distributive mixing characteristics. So the 180◦ phase angle case fared best in

terms of both cluster distribution index and interchamber material transfer, and 45◦

phase angle was the worst. In addition, the distributive mixing performance of 90◦

phase angle was slightly behind the 180◦ phase angle case.

121
The computational model, developed here, has been expanded to apply in 3D sim-

ulations for the rest of the research topics, i.e., analysis of friction speed ratios and

rotor designs.

2. Speed Ratios: (a) Dispersive mixing, which is very vital in breaking up the ingredi-

ent agglomerate and droplets, has been evaluated using statistical tools such as the

jPDF of mixing index and shear rate, and maximum shear stress. The jPDF signi-

fies the plug flow and elongational flow in conjunction with shear rate. From the

jPDF of mixing index and shear rate, it can be concluded that the speed ratio of 1.5

does provide with the best elongational flow components, but they are at very low

shear rates. Hence, these conditions may not be necessarily good in breaking the

agglomerates. This case also has the highest plug flow contribution. Plug flow is un-

desirable for mixing, since, higher the plug flow, the worse is the mixing due to more

material getting attached to the rotor without undergoing any stress. The jPDF also

indicates lower elongational flow for the speed ratio of 1.125, but also lower plug

flow. In addition, the speed ratio of 1.125 demonstrated higher percentage of parti-

cles undergoing maximum shear stress in comparison with other speed ratios, thus

further enhancing ingredient breakup. It should be noted that the maximum shear

stress analysis was based on the data from one revolution of the left rotor, and the

assumption in the current study is that, it is sufficient for the conclusions presented

here. Based on these results, one could summarize that 1.125 and 1.5 exhibited the

best and worst dispersive mixing capabilities, respectively, in this study.

122
(b) Furthermore, distributive mixing was qualitatively evaluated using axial distri-

bution, interchamber exchange of particles, transfer of particles between upper and

lower halves, each of which is critical in achieving homogeneity throughout the do-

main. Quantitatively, temporal evolution of axial distribution, cluster distribution

index and segregation scale were investigated for obtaining very insightful informa-

tion about the distributive mixing capability. Cluster distribution index for instance,

evaluates the deviation of the calculated distribution of the particles from the ideal

distribution. Based on the cluster distribution index, the speed ratio of 1.125 was the

closest to the ideal distribution at the end of the 20 revolutions, whereas the speed

ratio of 1.5 was the farthest. Axial distribution analysis also predicted a similar pat-

tern. In addition, the segregation scale showed that the speed ratio of 1.5 levels out

at a much higher value compared to the other two speed ratios, indicating that there

were still a lot of segregated particles at the end of the simulation. The speed ratio

of 1.125 showed a superior performance in that aspect, while the even speed perfor-

mance fell between 1.125 and 1.5. Here again, 1.125 and 1.5 exhibited the best and

worst distributive mixing capabilities, respectively, in this study.

3. Rotor Designs: Flow patterns and flow rates across various cross-sections were an-

alyzed to understand the material movement particularly due to wing configuration

and motion.

(a) Joint probability density plot of mixing index and shear rate was used for the

quantitative evaluation of dispersive mixing since both elongation and shear are crit-

123
ical for the breakdown of agglomerates. The 4-wing A rotor showed shear flow

mixing index regions (i.e. 0.5 < λMZ < 0.6) having higher shear rates than any of

the other two rotors.

(b) Distributive mixing was investigated here qualitatively using particle distribution

snapshots, and quantitatively using different particle statistics such as cluster distri-

bution index and length of stretch. Cluster distribution index evaluates the difference

between predicted distribution from simulation and ideal or perfect distribution. As

the cluster distribution index decreased with time/revolutions, the particle distribu-

tion got closer to an ideal one. Length of stretch increased with time as particle

moved further from neighboring particles and gets distributed. Oscillatory behavior

of the length of stretch curve implied stretching and unstretching or folding of the

fluid element as the fluid moved from one chamber cavity to the other and in axial di-

rection. The 4-wing A rotor not only reached the lowest value of cluster distribution

index the fastest, but also maintained a lower value compared to other two rotors.

The length of stretch of 4-wing A was also higher compared to the other two rotors.

All of these indicated better dispersive and distributive mixing capabilities of the

4-wing A rotor compared to the other two rotors. From a design perspective, this

was mainly due to two reasons: (i) Compared to the 2-wing, the 4-wing A rotor had

two additional short wings, which randomized and intensified the material movement

further. (ii) Compared to the 4-wing B rotor, the 4-wing A rotor had the two long

wings on the opposite side of each rotor, as opposed to the same side, and hence,

124
resulted in a higher shear rate and a more back-and-forth material movement among

the chambers.

While calculations presented here resulted in valuable information on the effects

of rotor design on mixing performance, several factors were not considered. For instance,

opportunity remains in the field for further improvement by extending the calculations to

non-isothermal through the additional consideration of viscous heating. Other factors such

as slip boundary conditions [74] and ram pressure also need further investigations. In

future, attempts can be taken to incorporate more of these realistic operating conditions in

order to improve the assessment and analysis of such mixing equipment, thereby providing

a better understanding of the various processes and also paving the way for more improved

designs.

125
BIBLIOGRAPHY

[1] Nortey, N. O., 1988. Internal batch mixing machines with non-intermeshing rotors
of increased performance, May 17. US Patent 4,744,668.

[2] Pointon, J. E., 1915. Machine for preparing rubber., May 4. US Patent 1,138,410.

[3] Banbury, F. H., 1916. Machine for treating rubber and other heavy plastic material.,
Oct. 3. US Patent 1,200,070.

[4] Comper, L. F., and Tyson, D. Z., 1966. Rubber mixer, Jan. 25. US Patent 3,230,581.

[5] Wiedmann, W., and Schmid, H.-M., 1980. Mixing apparatus for kneading of plastic
substances, Nov. 18. US Patent 4,234,259.

[6] Nortey, N. O., 1987. Two-wing non-intermeshing rotors of increased performance


for use in internal batch mixing machines, Dec. 22. US Patent 4,714,350.

[7] Takakura, K., Ureshino, K., Yamada, N., and Kurokawa, Y., 1998. Hermetically
closed kneading apparatus, Aug. 11. US Patent 5,791,776.

[8] Inoue, K., Ureshino, K., Yamada, N., Takakura, K., and Kurokawa, Y., 1999. En-
closed kneading apparatus, Nov. 16. US Patent 5,984,516.

[9] Inoue, K., Yamada, N., Hagiwara, K., and Nakano, H., 2010. Batch mixer and
mixing rotor used in the same, Dec. 21. US Patent 7,854,542.

[10] Valsamis, L., Borzenski, F., Wagner, R., Rapetski, W., and Baurmeister, H., 2002.
Four wing, non-intermeshing rotors for synchronous drive to provide improved dis-
persive and distributive mixing in internal batch mixers, Dec. 17. US Patent 6,494,607.

[11] Limper, A., Keuter, H., Rinker, M., and Berkemeier, D., 2009. Internal mixer for
kneading plastic materials, July 7. US Patent 7,556,420.

126
[12] Yoshida, N., Uemura, M., Nakano, H., Hagiwara, K., Inoue, K., Nishida, M., and
Fukutani, K., 2015. Kneading rotor, batch kneader and method of kneading materi-
als, Jan. 6. US Patent 8,926,166.

[13] Rupert, T. C., 1935. Rubber mixing or preparing machine, Sept. 24. US Patent
2,015,618.

[14] Johnson, F., Homann, H., Rother, H., and Weckerle, G., 1986. Mixing machine,
Feb. 5. EP Patent 0,170,397A1.

[15] Passoni, G. C., 1988. Closed parallel-rotor mixer with adjustable interaxial separa-
tion, Oct. 4. US Patent 4,775,240.

[16] Limper, A., 2012. Mixing of rubber compounds. Carl Hanser Verlag GmbH Co KG.

[17] Freakley, P., and Patel, S., 1985. “Internal mixing: A practical investigation of
the flow and temperature profiles during a mixing cycle”. Rubber chemistry and
technology, 58(4), pp. 751–773.

[18] Tadmor, Z., and Gogos, C. G., 2013. Principles of polymer processing. John Wiley
& Sons.

[19] Liu, J., Li, F., Zhang, L., and Yang, H., 2015. “Numerical simulation of flow of
rubber compounds in partially filled internal mixer”. Journal of Applied Polymer
Science.

[20] Manas-Zloczower, I., 2012. Mixing and compounding of polymers: theory and
practice. Carl Hanser Verlag GmbH Co KG.

[21] Freakley, P. K., 2012. Rubber processing and production organization. Springer
Science & Business Media.

[22] Wildi, R. H., and Maier, C., 1998. Understanding compounding. Hanser Verlag.

[23] Tokita, N., and White, J. L., 1966. “Milling behavior of gum elastomers: Experiment
and theory”. Journal of Applied Polymer Science, 10(7), pp. 1011–1026.

[24] White, J. L., 1992. “Development of internal-mixer technology for the rubber in-
dustry”. Rubber chemistry and technology, 65(3), pp. 527–579.

[25] Banbury, F. H., 1917. Machine for treating rubber and other heavy plastic material.,
May 22. US Patent 1,227,522.

127
[26] Nortey, N. O., 1989. Optimized four-wing, non-intermeshing rotors for synchronous
drive at optimum phase relation in internal batch mixing machines, May 30. US
Patent 4,834,543.

[27] Harada, J., and Nishigai, K., 1989. Rubber-like material kneading apparatus, Oct. 3.
US Patent 4,871,259.

[28] Nakajima, N., 2000. Science and practice of rubber mixing. iSmithers Rapra Pub-
lishing.

[29] Manas-Zloczower, I., Nir, A., and Tadmor, Z., 1984. “Dispersive mixing in rubber
and plastics”. Rubber chemistry and technology, 57(3), pp. 583–620.

[30] Vyakaranam, K. V., Ashokan, B. K., and Kokini, J. L., 2012. “Evaluation of effect
of paddle element stagger angle on the local velocity profiles in a twin-screw con-
tinuous mixer with viscous flow using finite element method simulations”. Journal
of Food Engineering, 108(4), pp. 585–599.

[31] Collin, V., Peuvrel-Disdier, E., Alsteens, B., Legat, V., Avalosse, T., Otto, S., and
Metwally, H., 2006. “Numerical and experimental study of dispersive mixing of
agglomerates”. In Society of Plastics Engineers Annual Technical Conference 2006,
ANTEC 2006, Vol. 2, Society of Plastics Engineers, pp. Pages–908.

[32] Wang, W., and Manas-Zloczower, I., 2001. “Temporal distributions: The basis for
the development of mixing indexes for scale-up of polymer processing equipment”.
Polymer Engineering & Science, 41(6), pp. 1068–1077.

[33] Yang, H.-H., and Manas-Zloczower, I., 1994. “Analysis of mixing performance in a
vic mixer”. International Polymer Processing, 9(4), pp. 291–302.

[34] Nikiel, L., Mason, R., and Wampler, W., 2016. “Advances in filler dispersion mea-
surements”. Rubber Chemistry and Technology, 89(1), pp. 142–153.

[35] Manas-Zloczower, I., 1997. “Analysis of mixing in polymer processing equipment”.


Rheology Bulletin, 66(1), pp. 5–8.

[36] Connelly, R. K., and Kokini, J. L., 2007. “Examination of the mixing ability of
single and twin screw mixers using 2d finite element method simulation with particle
tracking”. Journal of food engineering, 79(3), pp. 956–969.

[37] Wong, T. H., and Manas-Zloczower, I., 1994. “Two-dimensional dynamic study
of the distributive mixing in an internal mixer”. International Polymer Processing,
9(1), pp. 3–10.
128
[38] Cheng, H., and Manas-Zloczower, I., 1998. “Distributive mixing in conveying ele-
ments of a zsk-53 co-rotating twin screw extruder”. Polymer Engineering & Science,
38(6), pp. 926–935.

[39] Manas-Zloczower, I., and Cheng, H., 1996. “Analysis of mixing efficiency in poly-
mer processing equipment”. In Macromolecular symposia, Vol. 112, Wiley Online
Library, pp. 77–84.

[40] Freakley, P., and Idris, W. W., 1979. “Visualization of flow during the processing of
rubber in an internal mixer”. Rubber Chemistry and Technology, 52(1), pp. 134–145.

[41] Freakley, P., and Patel, S., 1987. “Internal mixing: A practical investigation of
the influence of intermeshing rotor configuration and operating variables on mix-
ing characteristics and flow dynamics”. Polymer Engineering & Science, 27(18),
pp. 1358–1370.

[42] Min, K., and White, J. L., 1985. “Flow visualization of the motions of elastomers
and molten plastics in an internal mixer”. Rubber chemistry and technology, 58(5),
pp. 1024–1037.

[43] Min, K., and White, J. L., 1987. “Flow visualization investigations of the addition
of carbon black and oil to elastomers in an internal mixer”. Rubber chemistry and
technology, 60(2), pp. 361–380.

[44] Min, K., 1987. “Flow visualization of blending of elastomers and thermoplastics in
an internal mixer”. Advances in Polymer Technology, 7(3), pp. 243–257.

[45] Morikawa, A., Min, K., and White, J. L., 1988. “Flow visualization of internal
mixer factory cycles for rubber compounding”. Advances in Polymer Technology,
8(4), pp. 383–405.

[46] Morikawa, A., Min, K., and White, J., 1989. “Flow visualization of the rubber
compounding cycle in an internal mixer based on elastomer blends”. International
Polymer Processing, 4(1), pp. 23–31.

[47] Kim, J. K., and White, J. L., 1989. “An experimental and theoretical study of star-
vation effects on flow and mixing elastomers in an internal mixer”. Nihon Reoroji
Gakkaishi (Journal of the Society of Rheology, Japan), 17(4), pp. 203–210.

[48] Kawanishi, K., and Yagii, K., 1990. “Flow analysis in an internal mixer: Part ii: Es-
timation of mixing efficiency by batch homogenization time”. International Polymer
Processing, 5(3), pp. 173–177.

129
[49] Kawanishi, K., Yagii, K., Obata, Y., and Kimura, S., 1991. “Relationship between
rotor designs in an internal mixer and physical properties of mixed rubber com-
pounds”. International Polymer Processing, 6(2), pp. 111–120.

[50] Kawanishi, K., Yagii, K., Obata, Y., and Kimura, S., 1991. “Scale-up effect in
internal mixers”. International Polymer Processing, 6(4), pp. 279–289.

[51] Cotten, G. R., 1984. “Mixing of carbon black with rubber i. measurement of dis-
persion rate by changes in mixing torque”. Rubber chemistry and technology, 57(1),
pp. 118–133.

[52] Griffith, R., Kannabiran, R., and Tomlinson, G., 1987. “Rubber flow in an internal
mixer”. Rubber chemistry and technology, 60(1), pp. 111–124.

[53] Menges, G., and Grajewski, F., 1988. “Process analysis of a laboratory internal
mixer”. International Polymer Processing, 3(2), pp. 74–78.

[54] Leblanc, J. L., and Lionnet, R., 1992. “Determining the components of mixing en-
ergy when preparing rubber compounds in instrumented internal mixers”. Polymer
Engineering & Science, 32(15), pp. 989–997.

[55] Toh, M., Gondoh, T., Mori, T., and Mishima, M., 2005. “Mixing characteristics of
an internal mixer: Uniformity of mixed rubber”. Journal of applied polymer science,
95(1), pp. 166–172.

[56] Kaewsakul, W., Sahakaro, K., Dierkes, W., and Noordermeer, J., 2012. “Optimiza-
tion of mixing conditions for silica-reinforced natural rubber tire tread compounds”.
Rubber chemistry and technology, 85(2), pp. 277–294.

[57] Ghari, H. S., Arani, A. J., and Shakouri, Z., 2013. “Mixing sequence in natural
rubber containing organoclay and nano-calcium carbonate ternary hybrid nanocom-
posites”. Rubber Chemistry and Technology, 86(2), pp. 330–341.

[58] Kiparissides, C., and Vlachopoulos, J., 1976. “Finite element analysis of calender-
ing”. Polymer Engineering & Science, 16(10), pp. 712–719.

[59] White, J. L., Kim, J.-K., Szydlowski, W., and Min, K., 1988. “Simulation of flow
in compounding machinery: Internal mixers and modular corotating intermeshing
twin-screw extruders”. Polymer Composites, 9(5), pp. 368–377.

[60] Kim, J., White, J., Min, K., and Szydlowski, W., 1989. “Simulation of flow and
mixing in an internal mixer”. International Polymer Processing, 4(1), pp. 9–15.

130
[61] Kim, J., and White, J., 1991. “Non-newtonian and non-isothermal modeling of 3d-
flow in an internal mixer”. International Polymer Processing, 6(2), pp. 103–110.

[62] Yagii, K., and Kawanishi, K., 1990. “Flow analysis in an internal mixer: Part i:
Application of finite element analysis”. International Polymer Processing, 5(3),
pp. 164–172.

[63] Manas-Zloczower, I., and Feke, D., 1989. “Analysis of agglomerate rupture in linear
flow fields”. International polymer processing, 4(1), pp. 3–8.

[64] Cheng, J.-J., and Manas-Zloczower, I., 1989. “Hydrodynamic analysis of a banbury
mixer 2-d flow simulations for the entire mixing chamber”. Polymer Engineering &
Science, 29(15), pp. 1059–1065.

[65] Cheng, J., and Manas-Zloczower, I., 1990. “Flow field characterization in a banbury
mixer”. International Polymer Processing, 5(3), pp. 178–183.

[66] Yao, C.-H., and Manas-Zloczower, I., 1997. “Influence of design on mixing effi-
ciency in a variable intermeshing clearance mixer”. International Polymer Process-
ing, 12(2), pp. 92–103.

[67] Yao, C.-H., and Manas-Zloczower, I., 1998. “Distributive mixing in an axial dis-
charge continous mixer”. International Polymer Processing, 13(4), pp. 334–341.

[68] Teverovskiy, M., Manas-Zloczower, I., Elemans, P., and Rekers, G., 2000. “Numer-
ical simulations and experiments in a double-couette flow geometry”. International
Polymer Processing, 15(3), pp. 242–254.

[69] Wang, W., Manas-Zloczower, I., and Kaufman, M., 2001. “Characterization of dis-
tributive mixing in polymer processing equipment using renyi entropies”. Interna-
tional Polymer Processing, 16(4), pp. 315–322.

[70] Wang, W., Manas-Zloczower, I., and Kaufman, M., 2003. “Entropic characterization
of distributive mixing in polymer processing equipment”. AIChE journal, 49(7),
pp. 1637–1644.

[71] Yao, C.-H., and Manas-Zloczower, I., 1996. “Study of mixing efficiency in roll-
mills”. Polymer Engineering & Science, 36(3), pp. 305–310.

[72] Gramann, P. J., and Osswald, T., 1992. “Simulating polymer mixing processes using
the boundary element method”. International Polymer Processing, 7(4), pp. 303–
313.

131
[73] Hutchinson, B., Rios, A., and Osswald, T., 1999. “Modeling the distributive mixing
in an internal batch mixer”. International Polymer Processing, 14(4), pp. 315–321.

[74] Alsteens, B., Avalosse, T., Legat, V., Marchal, T., and Slachmuylders, E., 2003.
“Effect of the full-slip condition along rotors on the mixing efficiency of internal
mixers.”. In ANTEC 2003 Conference Proceedings, pp. 173–177.

[75] Alsteens, B., Legat, V., and Avalosse, T., 2004. “Parametric study of the mixing effi-
ciency in a kneading block section of a twin-screw extruder”. International Polymer
Processing, 19(3), pp. 207–217.

[76] Salahudeen, S. A., Elleithy, R. H., AlOthman, O., and AlZahrani, S., 2011. “Com-
parative study of internal batch mixer such as cam, banbury and roller: Numerical
simulation and experimental verification”. Chemical Engineering Science, 66(12),
pp. 2502–2511.

[77] Salahudeen, S. A., AlOthman, O., Elleithy, R. H., Al-Zahrani, S., and Rahmat, A.
R. B., 2013. “Optimization of rotor speed based on stretching, efficiency, and vis-
cous heating in nonintermeshing internal batch mixer: simulation and experimental
verification”. Journal of Applied Polymer Science, 127(4), pp. 2739–2748.

[78] Dhanasekharan, K. M., and Kokini, J. L., 2003. “Design and scaling of wheat dough
extrusion by numerical simulation of flow and heat transfer”. Journal of Food Engi-
neering, 60(4), pp. 421–430.

[79] Connelly, R. K., and Kokini, J. L., 2004. “The effect of shear thinning and differ-
ential viscoelasticity on mixing in a model 2d mixer as determined using fem with
particle tracking”. Journal of Non-Newtonian Fluid Mechanics, 123(1), pp. 1–17.

[80] Rathod, M. L., and Kokini, J. L., 2013. “Effect of mixer geometry and operating
conditions on mixing efficiency of a non-newtonian fluid in a twin screw mixer”.
Journal of Food Engineering, 118(3), pp. 256–265.

[81] Nassehi, V., and Salemi, R., 1994. “Finite element modelling of non-isothermal vis-
cometric flows in rubber mixing”. International Polymer Processing, 9(3), pp. 199–
204.

[82] Nassehi, V., and Ghoreishy, M., 1997. “Simulation of free surface flow in partially
filled internal mixers”. International Polymer Processing, 12(4), pp. 346–353.

[83] Nassehi, V., and Ghoreishy, M., 2001. “Modeling of mixing in internal mixers with
long blade tips”. Advances in Polymer Technology, 20(2), pp. 132–145.

132
[84] Pokriefke, G., 2007. “Numerical simulation of viscous flow in a partially filled co-
rotating twin screw extruder”. International Polymer Processing, 22(1), pp. 61–72.

[85] Fukutani, K., Higashi, K., Yamada, S., and Yamane, Y., 2013. “Numerical study
on distributive mixing characteristics in a partially filled internal mixer”. In 184th
Technical Meeting of Rubber Division, ACS, no. 71, American Chemical Society.

[86] Wu, S., 1971. “Calculation of interfacial tension in polymer systems”. In Journal of
Polymer Science Part C: Polymer Symposia, Vol. 34, Wiley Online Library, pp. 19–
30.

[87] Wiedmann, W., and Schmid, H., 1982. “Optimization of rubber mixing in internal
mixers”. Rubber Chemistry and Technology, 55(2), pp. 363–381.

[88] Carreau, P. J., 1972. “Rheological equations from molecular network theories”.
Transactions of The Society of Rheology (1957-1977), 16(1), pp. 99–127.

[89] Yasuda, K., 1979. “Investigation of the analogies between viscometric and linear
viscoelastic properties of polystyrene fluids”. PhD thesis, Massachusetts Institute of
Technology.

[90] Escudier, M., Poole, R., Presti, F., Dales, C., Nouar, C., Desaubry, C., Graham, L.,
and Pullum, L., 2005. “Observations of asymmetrical flow behaviour in transitional
pipe flow of yield-stress and other shear-thinning liquids”. Journal of non-newtonian
fluid mechanics, 127(2), pp. 143–155.

[91] Patankar, S., 1980. Numerical heat transfer and fluid flow. CRC press.

[92] Muzaferija, S., Peric, M., Sames, P., and Schellin, T., 1998. “A two-fluid navier-
stokes solver to simulate water entry”. In Proceedings of the 22nd symposium on
naval hydrodynamics, Washington, DC, pp. 277–289.

[93] Brackbill, J., Kothe, D. B., and Zemach, C., 1992. “A continuum method for mod-
eling surface tension”. Journal of computational physics, 100(2), pp. 335–354.

[94] Vyakaranam, K. V., and Kokini, J. L., 2012. “Prediction of air bubble dispersion in
a viscous fluid in a twin-screw continuous mixer using fem simulations of dispersive
mixing”. Chemical Engineering Science, 84, pp. 303–314.

[95] Keuter, H., and Rinker, M., 2005. “Rotor optimization to improve mixing effi-
ciency.”. In European Rubber Research-Practical Improvements of the Mixing Pro-
cess: SATPRO, ROTOR and Dust Stop.

133
[96] Vyakaranam, K., Evans, M., Ashokan, B., and Kokini, J. L., 2009. “12 evaluation
of mixing and air bubble dispersion in viscous liquids using numerical simulations”.
Food Mixing: Principles and Applications, p. 253.

[97] Avalosse, T., Alsteens, B., and Legat, V., 2005. “Rotor shape design by numeri-
cal simulation: A new way to improve dispersive and distributive mixing in batch
mixer.”. In European Rubber Research-Practical Improvements of the Mixing Pro-
cess: SATPRO, ROTOR and Dust Stop.

[98] Danckwerts, P., 1952. “The definition and measurement of some characteristics of
mixtures”. Applied Scientific Research, Section A, 3(4), pp. 279–296.

[99] Danckwerts, P., 1953. “Theory of mixtures and mixing”. Research, 6, pp. 355–361.

[100] Ottino, J. M., 1989. The kinematics of mixing: stretching, chaos, and transport,
Vol. 3. Cambridge university press.

[101] Schuster, R. H., 2005. “Dispersion of fillers- a decisive factor for the processing
of rubber compounds and the performance of elastomers”. In European Rubber
Research-Practical Improvements of the Mixing Process: SATPRO, ROTOR and
Dust Stop.

[102] Emin, M., and Schuchmann, H., 2012. “Analysis of the dispersive mixing efficiency
in a twin-screw extrusion processing of starch based matrix”. Journal of food engi-
neering, 79(3), pp. 956–969.

134

You might also like