Sunteți pe pagina 1din 116

2016•2017

FACULTEIT INDUSTRIËLE INGENIEURSWETENSCHAPPEN


master in de industriële wetenschappen: chemie

Masterproef
Development and evaluation of an acrylic lubricant for PVC formulations.

Promotor :
Prof. dr. ir. Jozefien DE KEYZER

Promotor :
dr. SOFIE SANNEN

Gezamenlijke opleiding Universiteit Hasselt en KU Leuven


Shana Coomans
Scriptie ingediend tot het behalen van de graad van master in de industriële
wetenschappen: chemie
2016•2017
Faculteit Industriële
ingenieurswetenschappen
master in de industriële wetenschappen: chemie

Masterproef
Development and evaluation of an acrylic lubricant for
PVC formulations.

Promotor :
Prof. dr. ir. Jozefien DE KEYZER

Promotor :
dr. SOFIE SANNEN

Shana Coomans
Scriptie ingediend tot het behalen van de graad van master in de industriële
wetenschappen: chemie
Preface
As part of my education in chemical industrial engineering, I did an internship at the modifier division
of Kaneka Belgium. During this internship, I have focused on the development and evaluation of an
acrylic lubricant for PVC formulations.

I have had a wonderful and, most important, a great learning experience at Kaneka Belgium.
However, all this could not have been possible without a number of people who have helped me
during this research and my internship. First of all, I would like to thank my external promoter, Dr.
Sofie Sannen, very much. Thank you for always being there and freeing time for me when I needed
help regarding my research of my master’s thesis. I have learned a lot from your expertise and
experience and I would not have accomplished what I did without your help. In addition I would like
to thank all other colleagues for helping me when I needed it and sharing their experience and know-
how.

I would also like to thank my internal promotor, Prof. dr. ir. Jozefien De Keyzer, for the quick and
helpful feedback on the tasks for this master’s thesis. Also, what I have learned from your lessons in
plastic engineering has proven to be very useful during my research.

Last of all I would like to thank my parents. Without the help and support of my parents I would not
even have sat here and have written this master’s thesis. Therefore, I would like to thank them with
all my heart for giving me the chance to study and making it possible to build a future of my own.
Table of Content
Preface..................................................................................................................................................... 1
List of tables ............................................................................................................................................ 7
List of figures ........................................................................................................................................... 9
List of reactions ..................................................................................................................................... 13
Glossary ................................................................................................................................................. 15
Abstract ................................................................................................................................................. 17
Abstract in Dutch ................................................................................................................................... 19
INTRODUCTION .......................................................................................................................................... 21
1 Context .......................................................................................................................................... 21
2 Objectives ...................................................................................................................................... 21
3 Strategy ......................................................................................................................................... 22
PART 1: LITERATURE REVIEW ....................................................................................................................... 25
1 Polymerization............................................................................................................................... 25
1.1 Free radical chain polymerizations .............................................................................................................25
1.1.1 Initiation .................................................................................................................................................................... 25
1.1.2 Propagation ............................................................................................................................................................... 27
1.1.3 Termination ............................................................................................................................................................... 28
1.1.4 Chain transfer............................................................................................................................................................ 28
1.1.5 Inhibition and retardation ........................................................................................................................................ 31
1.2 Polymerization techniques .........................................................................................................................32
1.2.1 Suspension polymerization ...................................................................................................................................... 32
1.2.2 Emulsion polymerization .......................................................................................................................................... 34
1.2.3 Comparison of suspension polymerization with emulsion polymerization ........................................................... 38

2 Polyvinyl chloride .......................................................................................................................... 40


2.1 Additives ......................................................................................................................................................40
2.1.1 Fillers ......................................................................................................................................................................... 40
2.1.2 Colorants ................................................................................................................................................................... 41
2.1.3 Stabilizers .................................................................................................................................................................. 41
2.1.4 Plasticizers ................................................................................................................................................................. 42
2.1.5 Polymeric Modifiers .................................................................................................................................................. 43
2.1.6 Lubricants .................................................................................................................................................................. 43
2.2 Processing ....................................................................................................................................................43
2.2.1 Gelation or fusion of PVC ......................................................................................................................................... 44
2.2.2 Extrusion.................................................................................................................................................................... 45
2.2.3 Injection molding ...................................................................................................................................................... 47
2.2.4 Calendering ............................................................................................................................................................... 47
2.2.5 Blow molding ............................................................................................................................................................ 49

3 Lubricants ...................................................................................................................................... 50
3.1 Internal lubricants .......................................................................................................................................50
3.2 External lubricants.......................................................................................................................................51
3.3 Current lubricants .......................................................................................................................................52
3.4 Plate–out .....................................................................................................................................................52
3.5 Acrylic Lubricants ........................................................................................................................................53
PART 2: MATERIALS AND METHODS .............................................................................................................. 55
1 Materials........................................................................................................................................ 55
1.1 Materials for polymerization ......................................................................................................................55
1.2 Materials for evaluation ..............................................................................................................................55
2 Methods ........................................................................................................................................ 56
2.1 General set-up and procedure of the polymerization experiment ..........................................................56
2.1.1 General set-up .......................................................................................................................................................... 56
2.1.2 Polymerization experiment ...................................................................................................................................... 57
2.1.3 After treatment ......................................................................................................................................................... 59
2.1.4 Analysis of the samples ............................................................................................................................................ 59
2.2 Lubricating properties .................................................................................................................................63
2.2.1 Brabender gelation test ............................................................................................................................................ 63
2.2.2 Capillary rheometry .................................................................................................................................................. 65

PART 3: RESULTS AND DISCUSSION ............................................................................................................... 69


1 Development of the polymerization recipe .................................................................................. 69
1.1 Preliminary results ......................................................................................................................................69
1.2 Influence of reactor temperature ..............................................................................................................70
1.3 Influence of initiator....................................................................................................................................70
1.3.1 Influence of initiator type ......................................................................................................................................... 71
1.3.2 Influence of initiator concentration ......................................................................................................................... 72
1.4 Influence of emulsifier ................................................................................................................................74
1.5 Influence of chain transfer agent ...............................................................................................................78
1.6 After treatment ...........................................................................................................................................81
1.6.1 Development of the procedure ............................................................................................................................... 81
1.6.2 Properties of the powder ......................................................................................................................................... 82
1.7 Other ............................................................................................................................................................84
1.7.1 Monomer dosing rate ............................................................................................................................................... 84
1.7.2 Maron stability test................................................................................................................................................... 85
1.7.3 Molecular weight determination ............................................................................................................................. 85
1.7.4 Determination of the particle size ........................................................................................................................... 86
1.8 Conclusion ...................................................................................................................................................86
2 Evaluation of the acrylic lubricants ............................................................................................... 87
2.1 Brabender gelation .....................................................................................................................................87
2.1.1 Influence of dosing amount ..................................................................................................................................... 88
2.1.2 Comparison to other lubricants ............................................................................................................................... 91
2.1.3 Influence of molecular weight on the lubricating properties................................................................................. 92
2.1.4 Influence of the coagulation process....................................................................................................................... 93
2.1.5 Other experiments .................................................................................................................................................... 95
2.2 Shear viscosity measurements (Contifeed) ...............................................................................................96
2.2.1 Reproducibility .......................................................................................................................................................... 96
2.2.2 Influence of ‘standard’ lubricants on shear viscosity............................................................................................ 100
2.2.3 Influence of the acrylic polymer on shear viscosity .............................................................................................. 102
2.3 Conclusion ................................................................................................................................................ 103
CONCLUSION........................................................................................................................................... 105
Bibliography......................................................................................................................................... 107
Attachments ........................................................................................................................................ 111
List of tables
Table 1: Chemicals used during polymerizations with their function ................................................... 55
Table 2: Chemicals with their function which are used during coagulation of the latex ...................... 55
Table 3: Overview of the components with their function and concentration (PHR) for every charge
added during the polymerization experiment with X representing a variable concentration ............. 57
Table 4: Overview of the time (min) and tests performed on the samples during polymerization, with
SC the solid content, ƞsp the specific viscosity and GC gas chromatography ........................................ 58
Table 5: Temperature program (°C) for the determination of residual L on a GC with a HP-FFAP
column ................................................................................................................................................... 60
Table 6: Temperature program (°C) for the determination of residual M on GC with an AT-WAX
column ................................................................................................................................................... 61
Table 7: Retention times (min) of L and the internal standard using a HP-FFAP column and the
temperature profiles described in Table 5 ............................................................................................ 61
Table 8: Retention time (min) of M and the internal standard using an AT-WAX COLUMN and the
temperature profiles described in Table 6 ............................................................................................ 61
Table 9: Temperature profile (°C) used during DSC measurements ..................................................... 62
Table 10: An overview of the used shear rates (s-1) for each run ......................................................... 67
Table 11: Different components and their amounts (PHR) in recipe AcrLub6...................................... 69
Table 12: Specific viscosity (-) results of the polymerizations at different temperatures (°C) for
AcrLub6 and 10B_1 ............................................................................................................................... 70
Table 13: Specific viscosity results (-) of AcrLub9 with the type and amount of initiator (PHR) for
AcrLub9_1 and 9_2................................................................................................................................ 71
Table 14: The actual and theoretical conversion rates (PHR/h) for AcrLub9_1 and 9_2 with the
corresponding initiator and monomer dosing time (min) .................................................................... 71
Table 15: The obtained specific viscosities (-) with different ways of F addition for AcrLub6 and 10A_1
............................................................................................................................................................... 72
Table 16: Results of the specific viscosity measurements and the corresponding concentration of F in
the initiation charge for AcrLub10A_1 and 10A_2 ................................................................................ 72
Table 17: The obtained specific viscosities (-) with different amounts (PHR) of P4 for AcrLub11_1 and
11_2 ....................................................................................................................................................... 73
Table 18: The obtained specific viscosities by using a combination of the redox initiator P4 and A81
(PHR) for AcrLub12_2, 13_1 and 13_2 .................................................................................................. 73
Table 19: The actual and theoretical conversion rates (PHR/h) for polymerization reactions
AcrLub12_2, AcrLub13_1 and AcrLub13_2 with their corresponding dosing times (min) and used
initiator system...................................................................................................................................... 74
Table 20: Influence of the emulsifier type and concentration (PHR) on the specific viscosity (-) and
thus the molecular weight .................................................................................................................... 75
Table 21: The actual and theoretical conversion rates (PHR/h) for AcrLub12_1 and AcrLub12_2 with
the corresponding initiator and monomer dosing time (min) for different emulsifier types............... 75
Table 22: Specific viscosity results (-) of the polymerizations with different concentration (PHR) of
chain transfer agent (M1) for AcrLub10B_1, 10B_2, 11_1 and 12_1.................................................... 79
Table 23: The actual and theoretical conversion rates (PHR/h) for polymerization reactions
AcrLub10B_1, AcrLub10B_2, AcrLub11_1 and AcrLub12_1 with their corresponding dosing times
(min) and added amounts of M1 (PHR) ................................................................................................ 79
Table 24: Specific viscosity results (-) of the polymerizations with different concentration (PHR) of
chain transfer agent (M1) conducted with EM15 as emulsifier............................................................ 80
Table 25: The actual and theoretical conversion rates (PHR/h) for polymerization reactions
AcrLub102_2, AcrLub15_1, AcrLub15_2 and AcrLub14_1 with their corresponding dosing times (min)
and added amount (PHR) of M1 ........................................................................................................... 81
Table 26: Various amounts (PHR) of EM12 and STA12 used during the coagulation process .............. 82
Table 27: Specific viscosity (-) results of the latex and the powder for AcrLub12_1, 12_2, 13_1, 15_1
and 15_2 ................................................................................................................................................ 83
Table 28: Influence of the monomer dosing time (min) on the specific viscosity (-) ............................ 84
Table 29: The resulting molecular weight (g/mol) according to GPC measurements and the
corresponding specific viscosity (-) ....................................................................................................... 85
Table 30: Composition of the standard PVC formulation: NWG ll ........................................................ 87
Table 31: The calculated pressures (bar) and the applied shear rate (s-1) for both shear viscosity
measurements on the standard NWG formulation along with the pressure differences and
percentage error ................................................................................................................................... 97
Table 32: The calculated pressures (bar) and the applied shear rate (s1) for both shear viscosity
measurements on the NWG formulation without Licowax E along with the pressure differences and
percentage error ................................................................................................................................... 97
Table 33: Comparison of the shear viscosities (Pa.s) for the same shear rate (s1) for the standard
NWG formulation along with the percentage error between the measurements ............................... 99
Table 34: Comparison of the shear viscosities (Pa.s) for the same shear rate(s-1) for the NWG
formulation without Licowax E along with the percentage error between the measurements ........ 100
Table 35: Percentage difference (%) of the shear viscosities of all NWG formulation (formulation 2-4)
relative to the standard NWG formulation (formulation 1) accompanied with the applied shear rate
(s-1) ....................................................................................................................................................... 101
Table 36: Percentage difference (%) of the shear viscosities of all NWG formulation (formulation 2-4)
relative to the standard NWG formulation (formulation 1) accompanied with the applied shear rate
(s-1) ....................................................................................................................................................... 103
Table 37: Reaction conditions and final product properties of all performed polymerization tests .. 111
List of figures
Figure 1: The effect of inhibitors and retarders on a polymerization reaction [2]. .............................. 32
Figure 2: Mechanisms for breakage of monomer droplets [24]. .......................................................... 34
Figure 3: Mechanisms for coalescence of monomer droplets [24]. ..................................................... 34
Figure 4: General representation of the emulsion polymerization process [2]. ................................... 36
Figure 5: Representation of homogenous nucleation of active micelles [21]. ..................................... 36
Figure 6: Surface tension (σ) and rate of reaction (Rp) in function of the conversion for emulsion
polymerization [22]. .............................................................................................................................. 37
Figure 7: Schematic representation of the different intervals during emulsion polymerization [21]. . 38
Figure 8: Schematic overview of different particle sizes in different polymerization techniques [22]. 39
Figure 9: Schematic representation of the fusion or gelation behavior of PVC [36]. .......................... 44
Figure 10: Representation of the PVC structure with A) the crystalline zones connected with tie
molecules and B) the primary particles which are formed out of the crystallites and tie molecules
[37]. ....................................................................................................................................................... 44
Figure 11: Representation of primary particles where A) displays the interaction between primary
particles through free polymer molecules and B) represents the structure after cooling and the
formation of secondary crystallites [37]. .............................................................................................. 45
Figure 12: Schematic representation of an extrusion line with: A) The extruder, B) The die, C) The size
and cooling unit, D) The pull rolls and E) The cutter [26]...................................................................... 46
Figure 13: Representation of the extruder [1]. ..................................................................................... 46
Figure 14: More detailed representation of an extrusion screw with: A) The shank, B) The key, D) The
helix angle, E) The root, F) The channel depth, G) The flight, H) The channel, I) The pitch, J) The
diameter and K) The tip [26]. ................................................................................................................ 47
Figure 15: Simplified representation of an injection molding machine with: A) the clamping unit, B)
the mold or die C) the plasticizing unit, D) the control unit and E) the temperature control unit [35].
............................................................................................................................................................... 47
Figure 16:Representation of a calendering line with: a) Winder and edge cutter; b) Cooling rolls; c)
Four-roll calender, F type; d) Extruder; e) Mixing roll mill [35]............................................................. 48
Figure 17: Different configurations of calender rolls with A) I-type; B) F-type; C) L-type and D) Z-type
calendars [35]. ....................................................................................................................................... 48
Figure 18: Representation of the cross section of the built up and created currents with: A) Kneading
current; B) Feeding current; C) Outlet current [35]. ............................................................................. 48
Figure 19: Representation of extrusion blow molding line where: A) the parison extrusion; B)
positioning of the blow mold; C) positioning in the blow station; D) Shaping and cooling; E) Removing
of the object from the mold [35]. ......................................................................................................... 49
Figure 20: Representation of internal and external lubrication in PVC processing [3]. ........................ 51
Figure 21: The influence of A) internal and B) external lubrication on the speed of the flow front of
the polymer melt [39]. .......................................................................................................................... 51
Figure 22: General set-up of the polymerization reactions with 1) the water bath, 2) the heaters, 3)
the reactor, 4) the stirrer, 5) the stirrer driver, 6) the baffles, 7) a thermocouple, 8) the temperature
recorder, 9) the reflux cooler, 10) the monomer vessel, 11) the magnetic stirrer, 12) the nitrogen
supply, 13) the pump for the monomer emulsion, 14) the vessel for the second component and 15)
the pump for the second component. .................................................................................................. 57
Figure 23: General set-up for the Brabender gelation tests with A) the completely assembled
equipment and B) the different components of the disassembled chamber[36]. ............................... 63
Figure 24: Typical Brabender fusion plastogram with (1) torque (Nm) versus time (min) and (2)
temperature (°C) versus time (min) ...................................................................................................... 64
Figure 25: Typical Brabender torque (Nm) versus temperature (°C) fusion plastogram ...................... 65
Figure 26: General set-up of the capillary rheometer in combination with the extruder .................... 65
Figure 27: More detailed representation of the piston and capillary of the Göttfert RG25................. 66
Figure 28: Influence of the emulsifier on the volume particle size distribution (µm) for AcrLub12_1
(red) and AcrLub12_2 (green) ............................................................................................................... 76
Figure 29: Influence of the emulsifier on the number particle size distribution (µm) for AcrLub12_1
(red) and AcrLub12_2 (green) measured with the Microtrac S350 ...................................................... 76
Figure 30: Influence of the emulsifier on the intensity particle size distribution (µm) for AcrLub12_1
(red) and AcrLub12_2 (green) measured with the Nanotrac Wave ll ................................................... 77
Figure 31: Influence of the emulsifier on the number particle size distribution (µm) for AcrLub12_1
(red) and AcrLub12_2 (green) measured with the Nanotrac Wave ll ................................................... 77
Figure 32: SEM picture of AcrLub12_2 which is 10 000 times magnified ............................................. 78
Figure 33: SEM picture of AcrLub12_1 which is 10 000 times magnified ............................................. 78
Figure 34: Specific viscosity (-) as function of the M1 concentration (PHR) for the recipes with EM17.
............................................................................................................................................................... 79
Figure 35: The specific viscosity (-) as function of the concentration (PHR) of M1 for the recipes with
EM15 ..................................................................................................................................................... 80
Figure 36: Volume particle size distribution (µm) of the obtained powder of AcrLub12_1 (red),
AcrLub12_2 (green), AcrLub13_1 (yellow), AcrLub15_1 (blue) and AcrLub15_1 (gray) ....................... 83
Figure 37: Number particle size distribution (µm) of the obtained powder of AcrLub12_1 (red),
AcrLub12_2 (green), AcrLub13_1 (yellow), AcrLub15_1 (blue) and AcrLub15_1 (gray) ....................... 84
Figure 38: SEM picture of AcrLub15_1 which is 1 000 times magnified ............................................... 86
Figure 39: SEM picture of AcrLub15_1 which is 1 000 times magnified ............................................... 86
Figure 40: Temperature (°C) versus torque (Nm) for test AcrLub15_1 (coagulated with Z PHR EM12)
added in a standard NWG formulation with 0.5, 1.0 and 5.0 phr. A chamber filling of 54 grams was
used ....................................................................................................................................................... 88
Figure 41: Temperature (°C) versus torque (Nm) for test AcrLub15_2 (coagulated with Z PHR EM12)
added in a standard NWG formulation with 0.5 and 5.0 phr. A chamber filling of 54 grams was used
............................................................................................................................................................... 88
Figure 42: Torque (Nm) versus temperature (°C) for test AcrLub15_1 (coagulated with Z PHR EM12),
of which 0.5, 1.0 and 5.0 PHR was added in the NWG formulation without Loxiol G-11 and Licowax E
(formulation 4), with a chamber filling of 54 grams ............................................................................. 89
Figure 43: Torque (Nm) versus temperature (°C) for test AcrLub15_1 (coagulated with Z PHR EM12),
of which 0.5 and 1.0 PHR was added in the NWG formulation without Licowax E (formulation 2), with
a chamber filling of 54 grams ................................................................................................................ 90
Figure 44: Torque (Nm) versus temperature(°C) for test AcrLub15_1 (coagulated with Z PHR EM12),
Kane Ace™ PA101 and the EVONIK sample of which 0.5 PHR was added in the NWG formulation
without Licowax E and Loxiol G-11 (formulation 4), with a chamber filling of 54 grams .................... 91
Figure 45: Torque (Nm) versus temperature (°C) for test AcrLub12_1, 12_2, 13_1 and 15_1 (all
coagulated with Z PHR EM12) of which 0.5 PHR was added in the NWG formulation without Licowax
E and Loxiol G-11 (formulation 4), with a chamber filling of 54 grams ................................................ 92
Figure 46: Torque (Nm) versus temperature (°C) for test AcrLub12_1, 12_2, 13_1 and 15_1 (all
coagulated with Z PHR EM12) of which 0.5 PHR was added in the NWG formulation without Licowax
E (formulation 2), with a chamber filling of 54 grams........................................................................... 93
Figure 47: Torque (Nm) versus temperature (°C) for test AcrLub15_1 (coagulated with Z and 3.33Z
PHR EM12) of which 0.5 PHR was added in the NWG formulation without Licowax E and Loxiol G-11
(formulation 4), with a chamber filling of 54 grams ............................................................................. 94
Figure 48: Torque (Nm) versus temperature (°C) for test AcrLub15_1 (coagulated with Z PHR EM12
and Z PHR STA12) of which 0.5 PHR was added in the NWG formulation without Licowax E and Loxiol
G-11 (formulation 4), with a chamber filling of 50 grams ..................................................................... 95
Figure 49: Torque (Nm) versus temperature(°C) for test AcrLub15_1 (coagulated with Z EM12) and
Kane AceTM PA310 was added in the NWG formulation without Licowax E and Loxiol G-11
(formulation 4), with a chamber filling of 54 grams ............................................................................ 96
Figure 50: Shear viscosity (Pa.s) as function of shear rate (s-1) for the repeated measurements on the
standard NWG formulation, including data points fitted with the Yasuda model ............................... 98
Figure 51: Shear viscosity (Pa.s) as function of shear rate (s-1) for the repeated measurements on the
NWG formulation without Licowax E (formulation 2), including data points fitted with the Yasuda
model..................................................................................................................................................... 99
Figure 52: Shear viscosity (Pa.s) versus shear rate (s-1) for the measured and fitted date for the
different NWG formulations with adjusted axes ................................................................................ 101
Figure 53: Shear viscosity (Pa.s) versus shear rate (s-1) for the measured and fitted data for a NWG
formulation without Licowax E (formulation 2) and 0.5 PHR AcrLub15_1 and for a NWG formulation
without the standard lubricants (formulation 4) with 0.5 PHR AcrLub15_1 and 1.0 PHR Kane AceTM
PA310 .................................................................................................................................................. 102
Figure 54: Temperature curve of AcrLub9_1 (orange), AcrLub9_2 (blue) and the bath (yellow) during
polymerization. Herein, important points are indicated .................................................................... 112
Figure 55: Shear viscosity (-) versus shear rate (s-1) for the measured and fitted date for the different
NWG formulations with adjusted axes ............................................................................................... 113
List of reactions
Reaction 1: Termination of growing polymer chains by combination [2]. ............................................ 22
Reaction 2: Termination of growing polymer chains by disproportionation [2]................................... 23
Reaction 3: Thermal decomposition of tert-butyl peroxide [10]. ......................................................... 25
Reaction 4: Thermal decomposition of 2,2'-Azobisisobutyronitrile (AIBN) [2]. .................................... 26
Reaction 5: Redox initiation of tert-butyl hydro peroxide under the impulse of Fe2+-ions [9]. ............ 26
Reaction 6: Redox initiation of tert-butyl hydro peroxide under the impulse of Fe3+-ions [9]. ............ 26
Reaction 7: Redox initiation with potassium or sodium peroxodisulphate and sodium metabisulfite
[2]. ......................................................................................................................................................... 27
Reaction 8: Possible ways of monomer addition to the primary free radical [10]. .............................. 27
Reaction 9: Possible ways of propagation [10]. .................................................................................... 27
Reaction 10: Different resonance structures of a polystyrene radical [19]. ......................................... 28
Reaction 11: Termination of growing polymer chains by combination [2]. .......................................... 28
Reaction 12: Termination of growing polymer chains by disproportionation [2]................................. 28
Reaction 13: General representation of a chain transfer reaction [10]. ............................................... 29
Reaction 14: Reinitiation after chain transfer [10]................................................................................ 29
Reaction 15: Intermolecular chain transfer reaction for polyethylene [10]. ........................................ 30
Reaction 16: Polymerization of methyl acrylate with intermolecular chain transfer [10].................... 31
Reaction 17: General reaction of the production of PVC [29]. ............................................................. 40
Reaction 18: Representation of the thermal degradation of PVC [1]. .................................................. 41
Glossary
CMC Critical micelle concentration
DSC Differential scanning calorimetry
EBS Ethylenebisstearamide
GC Gas chromatography
GPC Gel permeation chromatography
SEC Size-exclusion chromatography
SEM Scanning electron microscopy
PHR Parts per hundred rubber
pPVC Plasticized polyvinyl chloride
PVC Polyvinyl chloride
PW Purified water
SC Solid content
uPVC Rigid polyvinyl chloride
Tg Glass transition temperature
THF Tetrahydrofuran
UV Ultra violet
VCM Vinyl chloride monomer
Abstract
Development and evaluation of an acrylic lubricant for PVC formulations.

One of Kaneka Belgium's activities is the development and production of additives for PVC in order to
improve its properties. Lubricants, for instance, are added to the PVC formulation to reduce friction
and melt viscosity in order to improve processing. However, current lubricants tend to display plate-
out, which is an uncontrolled deposition on machinery parts. Therefore, this master's thesis aimed to
develop a low molecular weight acrylic lubricant that is sufficiently compatible with the PVC matrix to
avoid plate-out.

The acrylic lubricant was synthesized via free radical suspension polymerization. Parameters of the
polymerization process, such as temperature, initiator type, initiator concentration, emulsifier type,
emulsifier concentration and addition of chain transfer agents were methodically adapted. As a
result, a recipe was achieved that produces low molecular weight acrylic polymers. From the
different polymerization tests, promising acrylic polymers were evaluated on their lubricating
properties by Brabender gelation and shear viscosity measurements. Results of these tests were
compared to the lubricating properties of standard lubricants.

A recipe to produce low molecular weight lubricating acrylic polymers was successfully developed.
The acrylic polymers displayed similar lubricating properties compared to current lubricants
according to Brabender gelation and shear viscosity measurements.
Abstract in Dutch
Ontwikkeling en evaluatie van een acrylic lubricant voor PVC formulaties.

Eén van de activiteiten van Kaneka Belgium is de ontwikkeling en productie van additieven voor PVC
om de eigenschappen te verbeteren. Zo worden bijvoorbeeld lubricants toegevoegd aan PVC
formulaties om de frictie en smeltvisocisteit te reduceren en zo de verwerking te verbeteren. De
huidige lubricants veroorzaken echter plate-out wat de ongecontroleerde afzetting is op
machineonderdelen. Daarom is in deze masterproef een laag moleculair acrylic lubricant ontwikkeld
dat voldoende compatibel is met de PVC matrix zodat plate-out wordt vermeden.

De acrylic lubricant werd gesynthetiseerd via radicalaire suspensie polymerisatie. Parameters van het
polymerisatieproces, zoals temperatuur, initiator type, initiator concentratie, emulgator type,
emulgator concentratie en toevoeging van chain transfer agent werden methodologisch aangepast.
Op deze manier werd een recept ontwikkeld dat acryl polymeren produceert met een laag moleculair
gewicht. De acryl polymeren met voldoende laag moleculair gewicht werden geëvalueerd met
behulp van Brabender gelatie en shear viscosity metingen. Deze resultaten werden vergeleken met
de eigenschappen van standaard lubricanten.

Er werd succesvol een recept ontwikkeld dat acrylic lubricants produceert met een laag moleculair
gewicht. Deze acrylic lubricants vertonen gelijkaardige lubrificerende eigenschappen met de huidige
lubricanten volgens Brabender gelatie en shear viscosity metingen.
INTRODUCTION

1 Context
Kaneka Belgium is a subsidiary company of Kaneka Corporation, which has its headquarters in Osaka
and Tokyo, Japan. Kaneka is an environmentally conscious and innovative company, playing a major
role globally in different fields such as functional plastics, life-science products, foodstuff products,
synthetic fibers,… Kaneka Belgium is established in Westerlo-Oevel. The production unit in which this
research project takes place is the modifier group. Here, additives for polyvinyl chloride (PVC) and
engineering plastics are produced. Applications can be found in impact, UV resistance and better
processability. Besides the modifier unit Kaneka Belgium consists of two other units: the liquid
polymer and the EPERAN foam division. Liquid polymers form the basis for adhesives and sealants.
They find applications in transportation, construction and do-it-yourself markets. EPERAN foams are
lightweight, environmentally friendly and recyclable, which makes them ideal for in shock-absorbing
packaging and automotive parts.

PVC is a polymer that is made out of vinyl chloride monomers. PVC is the second most produced
thermoplastic in the world after polyolefins and it has a wide scope of applications. PVC is used in
construction, wire and cable insulation, packaging, flooring,…. and can be processed using different
techniques such as extrusion, injection molding, calandering and blow molding [1][2][3].
One of the major problems with PVC formulations is its high melt viscosity, making the processing
difficult. As a consequence, high heat and power consumption is necessary in order to process the
polymer. Another disadvantage of PVC is that it sticks easily to the metal parts of the processing unit.
To avoid these problems, lubricants are added to PVC formulations. These lubricants influence how
PVC melts and flows during the processing [4]. The function of the lubricants can be expressed in two
different ways. On one hand, they lower the friction and adhesion during processing on the working
surface (external lubrication). On the other hand, they reduce friction by easing the movement of the
polymer chains in relation to the other polymer chains during processing. This has a positive
influence on the viscosity of the melt and heat build-up (internal lubrication) [1][3][4][5].
Current lubricants, such as polyethylene and propylene waxes, display some undesirable properties.
Plate-out is one of the most distinct problems during the processing of PVC. Plate-out is the
phenomenon that causes uncontrolled deposition on parts of the processing unit under influence of
lubricants. These lubricants, which are mainly hydrocarbons, cause plate-out by dissolving inorganic
additives and stabilizers in the PVC melt and act as a carrier for these additives. This complex of
lubricant and inorganic compounds can adsorb to parts of the processing unit and cause deposition.
As a consequence, final products will display striping and scoring, making them less appealing and
strong [6][7].

2 Objectives
The main goal of this research was to develop an acrylic external lubricant that solves the current
plate-out issue and improves final product performance such as gloss and surface finish. As a result,
standard polyethylene and propylene waxes can be omitted. For lubricants to practice their external
lubrication effect, they have to go to the outside of the PVC structure. Because the polyethylene and
propylene waxes have low compatibility with the PVC matrix, they will completely separate from the
matrix and form a film between the metal surface and the PVC melt [3]. This allows the melt to move
smoothly along the metal surface of the processing unit. But, due to the formation of the second
layer, these lubricants tend to express plate-out more easily, as previously mentioned.
The proposed acrylic polymer has a low molecular weight and a higher compatibility with PVC. The
low molecular weight polymer chains, on the on hand, causes the acrylic lubricant to go to the wall of
the processing unit under the influence of shear. Because of the higher compatibility, on the other
hand, the acrylic polymer will not separate completely from the PVC melt. As a result, the acrylic
copolymer lubricates the melt but does not display the plate-out problem.
The acrylic copolymer consists of a carbon backbone, having C12 to C18 side chains. These relatively
long side chains are required so that the lubricant will go to the outside of the PVC matrix. To obtain
this specific structure, a mixture of monomer M and L was polymerized. The resulting acrylic
polymers were evaluated based on their lubricating efficiency, with a Brabender gelation test and
capillary viscosity measurements. This research is successful when a lubricant is developed with the
same lubricating properties as the current resources for standard PVC formulations, but with which
the current plate-out problem can be omitted.

3 Strategy
The lubricant was developed using free radical polymerization. In this type of polymerization, the
monomer units are added to a growing polymeric chain via radical reactions in three steps. In a first
step, free radicals are produced from an initiator during the initiation. The chemicals which are used
for initiation can decompose easily under impulse of heat or by a reducing agent. Peroxides,
hydroperoxides and azo compounds are most frequently used as initiators. The initiation usually
takes place in two phases. In the first phase, free radicals are formed out of the initiator. Afterwards,
the free radicals are added to a monomer molecule. The initiated monomer can react further with
the rest of the present monomer and start the second step: the propagation. In this step the polymer
chain grows while maintaining the reactive radical at the end of the chain. The propagation step
proceeds until the free reactive end of the polymer chain is ended in a last step: the termination
[2][8][9][10]. The two most common mechanisms for termination are combination and
disproportionation. Combination occurs when two growing polymer chains combine, which results in
one large polymer chain (Reaction 1)[2][10].

Reaction 1: Termination of growing polymer chains by combination [2].

Disproportionation on the other hand occurs when a hydrogen atom from one of the growing chains
is abstracted by another. This kind of termination results in the formation of two different polymer
chains, enhancing low molecular weight (Reaction 2) [2][10].

22
Reaction 2: Termination of growing polymer chains by disproportionation [2].

To carry out the free radical polymerization, a suspension process was used in a (semi-) batch
reactor. Suspension polymerization is a type of dispersion or heterogeneous polymerization. This
means that liquid monomer is dispersed in a continuous phase with an amphiphilic compound, since
the monomer and polymer are insoluble in the continuous phase. Water was used as the continuous
phase in the polymerizations in which monomer droplets with a diameter between 10 and 100 µm
are dispersed. In order to avoid coalescence of the droplets, the reactor was agitated. To start free
radical polymerization, addition of an initiator to the reaction was necessary. Due to the insolubility
of the monomer in the aqueous phase, the initiator should be completely or partially soluble in the
monomer droplets to ensure polymerization takes place [2][9][11]. The initiator decomposed
thermally or via a reducing agent and started the polymerization reaction in the monomer droplets
according to the mechanism that was previously mentioned.

A recipe was developed to produce a low molecular weight lubricant. In a first attempt to lower the
molecular weight, the temperature was increased. By increasing the reaction temperature, the rate
of termination by disproportionation increases and so the rate of termination by combination
decreases [9][10]. Increasing the temperature also promotes the decomposition of the initiator
which enhances the radical concentration. An increase in radical concentration increases the
formation of the polymer chains and thus the chains should be shorter. Subsequently, an attempt
was made to rise the radical concentration even more by changing the initiator type, using different
combinations and concentrations [2][9][12]. The emulsifier type and concentration were also change
in order to increase the radical flux into the monomer drops.
In order to reach the target molecular weight, the addition of a chain transfer agents was necessary.
Chain transfer agents replace the activity of a growing polymer chain to another molecule, lowering
the molecular weight of the polymer. The growing polymer chain abstracts an atom (mostly a
hydrogen or halogen atom) from the chain transfer agent. As a result, a ‘dead’ polymer chain and a
new radical are formed, which can start a new chain [10][13]. The acrylic monomers may also
undergo chain transfer reactions. The extent to which the monomer undergoes the chain transfer
reaction is dependent on its transfer constant. The transfer constant (CTA) is the ratio between the
rate coefficient for chain transfer to the monomer (ktr), to the rate coefficient of propagation (kp). As
a result, a monomer with a large CTA will be more sensitive to chain transfer, enhancing a lower
molecular weight [2][10]. There is thus an optimum ratio between the used monomer acrylics, in
order to produce the lowest molecular weight.

During all polymerization reactions, conversion was followed using gas chromatography. Via gas
chromatography it is possible to detect the unreacted monomer, from which the conversion of the
polymerization reaction can be determined. This procedure was also used to determine the residual
monomer on the end of the polymerization and determining overall conversion.
The synthesized polymers were evaluated afterwards using different analytical methods. The
molecular weight was determined by measuring the specific viscosity. Out of the measured specific

23
viscosity the molecular weight was derived via an empirical relation [14]. In addition, GPC
measurements were conducted to obtain the molecular weight distribution.
Particle size distribution measurements were important for two different reasons. First, it was
measured because of the effect on the lubricating properties. Small polymer particles detangle more
easily into individual polymer chains under the influence of shear, meaning they will express their
lubricating effect more easily. Secondly, particle size distribution was important in order to verify the
droplet size during the polymerization. The particle size was important because it relates to the rate
of polymerization and thus to the final molecular weight. Measurements on particle size and particle
size distribution were based on the law of Lambert–Beer and Mie theory[15][16][17].
Final lubricating properties of potential lubricating acrylics were determined using Brabender tests
and capillary rheometry. The Brabender test is typical for PVC formulations to measure their fusion
or gelation behavior. Fusion of PVC is the process in which a resin grain is broken down into primary
particles and eventually forming a homogenous melt under the influence of shear, friction and heat.
In a Brabender Plastograph the PVC grains are masticated in a heated bench internal mixer. The PVC
formulation will exert a torque due to the masticating and ‘melting’ the PVC. The average torque
which was measured during the mixing of the PVC at constant temperature, was an indication for the
viscosity of the PVC melt and thus a value for lubrication [18]. A reduction in torque was observed
when adding an acrylic lubricant to the PVC formulation. Lubricating properties can also be
determined using capillary rheometry. The PVC melt is forced through the capillary at a constant
rate. Out of the applied flow and pressure drop over the capillary, shear viscosity was determined
[19]. Due to the lubricating effects of the acrylics, a reduction in shear viscosity was observed.

24
PART 1: LITERATURE REVIEW

1 Polymerization
Polymerization is the process where, hundreds to thousands of monomer molecules are connected
via chemical reaction. As a result of this connection, a macromolecule or polymer is formed
[9][2][20]. Polymers can be classified according to two different polymerization mechanisms. On one
hand, there is step growth polymerization which proceeds via consecutive reactions of the functional
groups of the monomers [2][20][19]. The second kind of polymerization is chain growth
polymerization. In this kind of polymerization the monomer is added to reactive centers onto a
growing polymer chain. This active center could either be anionic, cationic or radical
[2][20][19][8][10]. In this work only radical chain growth polymerization will be used and will
therefore be discussed more in depth in the following section.

1.1 Free radical chain polymerizations


Free radical polymerization is the most used type op chain polymerization. It is practiced for
polymerization of monomers which have a general structure like CH2=CR1R2. Where R1 and R2 are
substituent groups [10]. During the chain growth polymerization a pi bond is modified into a sigma
bond going from a sp2 bond is modified to a sp3 hybridization [8].
Making a polymer with free radical chain polymerization proceeds in three steps: initiation,
propagation an termination [9][20][19][10]. The course of the three steps of free radical chain
polymerization will be discussed more in depth in the following sections.

1.1.1 Initiation
The initiation of the polymerization takes place in two steps. First, free primary radicals are formed
from an initiator. These radicals are very reactive species because they contain an unpaired electron.
Therefore, the radicals will undergo a reaction to withdraw an electron from another molecule.
Because of this high reactivity, these free radicals will react with the monomer molecules in a second
step[20][10].
Formation of free radicals can proceed according to homolytic scission or redox initiation
[9][2][20][10]. Homolysis or homolytic scission of the initiator can occur by the application of heat,
which is called thermolysis. In order to use a compound as thermal initiator, the compound has to
have a dissociation energy in the range of 100 to 170 kJ mol-1. For dissociation energies out of this
range, dissociation will occur either too rapid or too slow. Because of this specific range, only
compounds with O-O-, S-S- or N-O-bonds are qualified to be used as a thermal initiator [2]. However,
peroxides are mostly used as a thermal initiator because of the instability and lack of provision of the
other two compounds. Tert-butyl peroxide is a thermal initiator which is frequently used [9][2][10].
Reaction 3 gives the thermal decomposition of this peroxide.

Reaction 3: Thermal decomposition of tert-butyl peroxide [10].

25
Also azo-compounds are frequently used as a thermal initiator. Although there is no weak bond
present in these molecules, the release of the highly stable molecular nitrogen is a sufficiently driving
force for decomposition into free radicals [2]. This release of molecular nitrogen is profitable for the
decomposition, but it can disturb polymerization reactions at high concentrations of the initiator [9].
An example of the thermal decomposition of 2,2’-Azobisisobutyronitrile (AIBN) is given in Reaction 4.

Reaction 4: Thermal decomposition of 2,2'-Azobisisobutyronitrile (AIBN) [2].

Homolytic scission is also possible by irradiation with light, defined as photolysis. Its advantage over
thermolysis is the formation of the radicals as soon as the initiator is exposed to the light source, and
ends as soon as the light source is removed. The use of photo-initiation is mostly used with heat
sensitive materials. However, due to the limited penetration of the light through materials, it is only
used in a limited number of applications [2][10]. Photo-initiators will not be used in this research and
therefore will not be discussed any further.
Redox initiation is a third way to generate free primary radicals. The radicals are formed through an
electron transfer between a reductant and oxidant. By using a redox system as initiator, radicals can
be produced at a sufficiently high speed at reduced temperatures in contrast to thermal initiation.
There is a variety of redox initiator systems available. Most frequently (hydro)peroxides are used as
an oxidizing agent, with metal ions as an reductant [9][2].

2+
Reaction 5: Redox initiation of tert-butyl hydro peroxide under the impulse of Fe -ions [9].

3+
Reaction 6: Redox initiation of tert-butyl hydro peroxide under the impulse of Fe -ions [9].

Often a component, such as sodium formaldehyde sulfoxylate (CH3NaO3S), is added to the reaction
mixture to accelerate the reduction of the Fe3+-ions to Fe2+-ions and recover the initial amount of
Fe2+-ions.
Another redox system, which is widely used as a source for free radicals, are persulfates in
combination with an inorganic reductant. For example, a redox system of potassium (K 2S2O8) or
sodium (Na2S2O8) peroxodisulphate with sodium metabisulfite (Na2S2O5) is often used (Reaction 7)
[2].

26
Reaction 7: Redox initiation with potassium or sodium peroxodisulphate and sodium metabisulfite [2].

In order for the free radicals to start polymerization, they have to attack the п-bond of the monomer
and form a new radical [2][19][8].
The addition of the monomer to the free radical can take place in two different ways. Reaction 8
gives the possible ways the free radical can add to the monomer[10].

Reaction 8: Possible ways of monomer addition to the primary free radical [10].

The addition will mainly proceed according to the first reaction because the attack of the radical R is
more sterically hindered by substituent X in the second reaction. Besides, the free radical which is
formed after addition to the monomer can be stabilized by mesomerism by substituent X in the first
reaction [2][10].

1.1.2 Propagation
In the propagation step, the polymer chain grows through addition of monomers to the active
center. This also results in the transformation of a п-bond into a σ-bond. Addition of one monomer
molecule to the growing chain takes about a millisecond, meaning thousands of monomer additions
can occur within a few seconds. The addition of the monomer to the polymer can proceed in two
different ways, head-to-tail addition or 1,3-placement (1) and head-to-head addition or 1,2-
placement (2). These possibilities are shown in Reaction 9 [2][10].

Reaction 9: Possible ways of propagation [10].

Propagation will occur mainly head-to-tail, because there is less steric hindrance present. Also,
substituent X may stabilize the formed free radical by the resonance effect in head-to-tail addition.
This is not possible with head-to-head placement. When mainly head-to-tail addition takes place, the
propagation is regioselective. However, this does not mean head-to-head addition will not occur at
all. When the substituent X is small and offers little resonance stabilization, high proportions of head-
to-head addition can be observed[2][10].
The rate at which the propagation proceeds is dependent on the reactivity of the terminal radical of
the polymer. Because of the high reactivity of the polymer-radical, reactivities of the monomers are

27
subservient. Resonance stabilization of the polymer-radical is an important factor in its reactivity.
For example, radicals of polystyrene are stabilized due to conjugation of unpaired electrons with the
unsaturated groups on the terminal carbons. Because of this high stabilization, polystyrene-radicals
are less reactive in comparison to for example (vinyl acetate)-radicals which are not able to stabilize
the radical. The different resonance structures for a polystyrene-radical are given in Reaction 10 [19].

Reaction 10: Different resonance structures of a polystyrene radical [19].

1.1.3 Termination
In the termination step, the reactive site of the growing polymer chain is ended. The two most
common mechanisms for termination are combination and disproportionation. Combination occurs
when two growing polymer chains combine, which results in one large polymer chain (Reaction
11)[2][10].

Reaction 11: Termination of growing polymer chains by combination [2].

Disproportionation on the other hand occurs when a hydrogen atom from one of the growing chains
is abstracted by another. This kind of termination results in the formation of two different dead
polymer chains. One of these terminated polymer chains has a saturated terminal group whereas the
other has an unsaturated terminal group. Due to the formation of two polymer chains,
disproportionation enhances low molecular weight chains (Reaction 12) [2][10].

Reaction 12: Termination of growing polymer chains by disproportionation [2].

The termination mechanism that occurs is dependent on the monomers which are used during the
polymerization. Most monomers tend to terminate more by combination because of high activation
energies necessary for termination by disproportionation. As a consequence, the rate of termination
by disproportionation increases with temperature. However, there are also compounds such as 1,1-
disubstituted olefins which tend to terminate more by disproportionation because of the steric
hindrance [2].

1.1.4 Chain transfer


Chain transfer is a third kind of termination of the growing polymer chain. This is the premature
termination of an active polymer chain. Reaction 13 gives a general representation of a chain transfer

28
reaction, in which ktrTA presents the rate constant and the TA chain transfer agent inducing the chain
transfer reaction [10].

Reaction 13: General representation of a chain transfer reaction [10].

Termination by chain transfer occurs when an atom (T), mostly a hydrogen or halogen atom, of a
present substance (TA) in the reaction medium is exchanged with the terminal radical of a growing
polymer chain. As a result, a dead polymer and a new free radical (A•) are generated. Most of these
free radicals are capable of reacting with a monomer to start the growth of a new chain. This is
displayed in Reaction 14 [2][19][10]. However, not every generated free radical out of chain transfer
is able of reinitiating the growth of a new chain, which is often called degenerative chain transferring
[19].

Reaction 14: Reinitiation after chain transfer [10].

All substances present in the reaction medium, such as solvents, monomers and initiators, can act as
a chain transfer agent (TA). Obviously, the ease at which the chain transfer takes place is dependent
on the strength of the bond between the exchanged atom of the chain transfer agent and the rest of
its molecule. In some cases, chain transfer agents are added to the reaction in order to intentionally
cause chain transfer. These chain transfer agents are mostly mercaptans and halogen compounds
due to their particularly labile atoms. This is desirable when polymers with low molecular weight are
aimed for [2][19][10][21].
Chain transfer to a polymer molecule can also occur, resulting in a branched polymer. When
intermolecular chain transfer occurs, mainly short-chain branches are formed. This phenomenon is
also known as back-biting. Reaction 15 gives the intermolecular chain transfer reaction for
polyethylene. This kind of chain transfer does not affect the molar mass of the polymer. However,
the branched structure does have an effect on the polymer properties[10].

29
Reaction 15: Intermolecular chain transfer reaction for polyethylene [10].

Chain transfer to a polymer can also occur intramolecular, resulting in a polymer with long branches.
This reaction takes place via the abstraction of an atom from a substituent group or a backbone
tertiary hydrogen atom. The latter reaction can occur in the polymerization of methyl acrylate, which
is shown in Reaction 16. As well as with intermolecular chain transfer, the molecular structure of the
polymer is changed. Also, due to the growth of the long chain branches, the molar mass distribution
of the polymer is broadened. Both the change in molecular structure and the broadening of the
molecular weight distribution can have a major effect on the polymer properties[10].

30
Reaction 16: Polymerization of methyl acrylate with intermolecular chain transfer [10].

1.1.5 Inhibition and retardation


Inhibition of a polymerization reaction occurs when substances react very fast with the free radicals
as the radicals are formed during the initiation and propagation As a result, new free radicals are
formed, but due to their stability they are mostly not able to initiate chain growth. This prevents the
polymerization to continue or even start until the inhibitor is consumed completely, resulting in an
induction period before the actual polymerization proceeds at a normal rate. Retardation on the
other hand, is the phenomenon which reduces the rate of the polymerization reaction. Retarders are
also species which react with the free radicals. However, the reaction is not as energetically as
reaction with inhibitors, making it possible for some of the free radicals to escape. Therefore, the
reaction does not have an induction period, but is slowed down the polymerization throughout its
course [2][10][19].
Figure 1 shows the effects of inhibitors and retarders on the conversion in function of time. Plot 1
represents the polymerization reaction in absence of an inhibitor or retarder. When an inhibitor is
introduced to the polymerization reaction, plot 2 is observed. This plot clearly shows that an
induction period is introduced before the polymerization reaction starts. Plot 3 shows the
polymerization reaction in the presence of a retarder. The reaction proceeds clearly at a much slower
rate. In some special cases, components display more complex behavior. They first act as a inhibitor
and are converted in a retarder after the induction period (plot 4) [2].

31
Figure 1: The effect of inhibitors and retarders on a polymerization reaction [2].

One of the best known and most important retarder and inhibitor is oxygen. Due to its
omnipresence, free-radical polymerizations need to be carried out under an inert atmosphere in
order to become reproducible results. Other compounds causing retardation and inhibition are
impurities, explaining irreproducible polymerization rates when impure monomers are used. On the
other hand, inhibitors are often used to prevent premature polymerization of the monomers during
transport or storage. These inhibitors are removed before polymerization of the monomers or an
excess of initiator is added to the reaction medium [2][10][10].

1.2 Polymerization techniques


Used techniques to carry out polymerization reactions can be classified into two groups,
homogenous and heterogeneous polymerizations. This classification is based on the initial state of
the mixture. Homogenous polymerization includes bulk polymerization and solution polymerization.
Bulk or mass polymerization is the simplest technique of polymerization. Here, the polymerization is
performed in pure monomer, which provides pure and solid polymers. A big disadvantage of this
method op polymerization is the high viscosity, making mixing and heat removal difficult. This
problem can be overcome by the addition of a solvent in which the monomer is soluble. This method
is called solution polymerization. However, impurities in the polymer and problems with solvent
removal may occur [2][20]. Heterogeneous polymerization consists of suspension polymerization and
emulsion polymerization. In these types of polymerization, monomers are dispersed in a solvent in
which they are not soluble. These polymerization processes are particularly suitable to deal with
thermal and viscosity problems[2]. Due to the importance of these techniques for this work, they will
be discussed more in depth in the following sections.

1.2.1 Suspension polymerization


Suspension polymerization is carried out by suspending the monomer in a continuous aqueous phase
with the aid of a suspending agent. Because the reaction occurs in the pure monomer drops, the
reaction kinetics are similar to those in bulk polymerization. Thanks to the monomer droplets, the
overall viscosity of the reaction medium in suspension polymerization is lower in comparison to bulk
polymerization, leading to a couple of advantages such as better heat removal and agitation.
However, the viscosity of the individual monomer drops increases when the conversion of the
monomers increases [19][22][11].
In general, suspension polymerization can be divided into two groups. In bead (suspension)
polymerization the formed polymer is soluble in the monomer drops. During polymerization, the
monomer drops are transformed from viscous drops into solid, smooth and spherical particles.

32
Whereas with powder (suspension) polymerization, the formed polymer is not soluble in the
monomer drops, causing precipitation of the polymer into irregular grains or particles[23][24].
Sometimes a third kind of suspension polymerization is distinguished, called mass suspension
polymerization. This kind of polymerization proceeds in two steps. In a first step a rubber is dissolved
in a liquid monomer mixture, which is then polymerized in a bulk process. After the mixture has
reached a conversion around 25% the reaction medium is transferred in a suspension reactor
containing water and suspending agent. In a second step, this mixture is polymerized via suspension
polymerization[23].
Most initiators used in suspension polymerization are soluble in the monomer drops and will start
the polymerization according to mechanism explained above in section 1.1 in the drops.
During this polymerization process, the dispersed monomer droplets are converted into spherical
polymer particles. Therefore, control of the monomer drop size is an important parameter in
suspension polymerization. These monomer droplets can have a diameter in the range of 10 to 100
µm and are kept in suspension trough vigorous stirring and the addition of surface active-agents. The
degree of agitation and the concentration of suspending agents are the two most important
parameters in the control of the particle size distribution of the final polymer particles[2][23][11][24].
As suspending agent, water-soluble organic macromolecules (e.g. poly (vinyl alcohol)) can be used.
These components adsorb on the monomer droplet to form a protective film or skin around the
droplet, preventing the approach and attachment of other droplets. These macromolecules also
reduce the interfacial tension which lowers the required energy for drop formation [2][23]
[11][25][26]. Insoluble and inorganic compounds (e.g. magnesium carbonate) can be used as well as
suspending agents. These components are able to stabilize the monomer droplets by the Pickering
effect. However, the mechanism of the Pickering effect of inorganic solids is still rather unclear. In
comparison to the organic macromolecules, these Pickering stabilizers have a few advantages. For
example, the concentrations needed for stabilization with inorganic components are much lower in
comparison to the concentrations needed with organic macromolecules. Additionally, the inorganic
components are much easier to remove in comparison to the organic macromolecules. The inorganic
components also decrease deposition of the polymer on the reactor wall, promoting the removal of
heat [2][23][11][26].
The drop and particle size distribution is controlled by breakage and coalescence rates, in which the
agitation and concentration of surface-active agents play an important role [22][24]. Breakage of the
droplets can proceed according to two mechanisms. A first mechanism is known as Thorough
breakage and occurs under the influence of viscous shear forces, splitting the monomer droplet into
two fluid lumps which are connected through a liquid thread. Eventually, this results in the formation
of two monomer drops with almost the exact same size and some smaller drops, corresponding with
the connecting thread (Figure 2(a))[24]. When a monomer droplet is suspended in a turbulent flow, it
is exposed to relative velocity and local pressure fluctuations. When the continuous phase and the
monomer droplets have nearly the same densities and viscosities, oscillating of the droplet surface
may occur. Subsequently, small droplets can be stripped out from the initial drop when the relative
velocity is close to relative velocity to make the droplet marginally unstable. This mechanism is called
erosive breakage, and is mostly the dominant mechanism in low-coalescence systems with
bimodality in particle size distribution (Figure 2(b)) [24][27].

33
Figure 2: Mechanisms for breakage of monomer droplets [24].

For the coalescence of monomer droplets, two mechanisms are suggested in the literature. In a first
mechanism, the coalescence is initially prevented by a thin layer of the continuous phase. Due to the
presence of attractive forces, the liquid film can be overcome resulting in coalescence of the two
monomer droplets (Figure 3(a)). However, when the kinetic energy of the oscillations of the original
droplet is larger than the adhesion energy, the drops will not coalesce [24]. In a second mechanism,
it is assumed that immediate coalescence of the monomer drops occurs when the turbulent energy
of collision is larger than the total drop surface energy. In other words, the velocity of the clashing
drops exceeds a critical value in velocity at the collision instant, resulting in one big drop (Figure
3(b))[24].

Figure 3: Mechanisms for coalescence of monomer droplets [24].

In perspective of coalescence and breakage, the suspension polymerization process can be divided in
three stages. At the start of the polymerization, breakage of the monomer drops is the dominant
mechanism, resulting in smaller drops. As the polymerization reaction proceeds, the viscosity of the
reaction medium is increased, and the so-called “sticky-stage” of the polymerization is reached at a
monomer conversion of 10-20%. In this stage, coalescence becomes more significant leading to an
increase in particle size. When the monomer conversion reaches 75-80%, the particles lose their
sticky behavior and the particles size distribution remains at a constant level. This level is called the
identification point for the particle size distribution. However, the particle size can slightly change
due to difference in density between the monomer and the polymer causing shrinkage of the
polymer particles [23][22][24][26].

1.2.2 Emulsion polymerization


In emulsion polymerization, a monomer which is partially soluble in water is dispersed in water with
the aid of an emulsifier as displayed in Figure 4. The used emulsifier is an important component in
emulsion polymerization because it strongly affects the final properties of the resulting polymer
latex. Emulsifiers consist of a hydrophilic and hydrophobic part and can either be anionic, cationic or
nonionic dependent on the electric charge of the hydrophobic part of the molecule. When the
solution with the emulsifier is strongly diluted, the emulsifier molecules will behave as individual

34
molecules or electrolytes. However, when the emulsifier concentration is increased, at a certain
point the emulsifier molecules form ordered molecular structures, known as micelles. This
concentration is the critical micelle concentration (CMC) and from this concentration, changes in
physical properties of the solution, such as viscosity, can be observed. The CMC ranges between
0.001–0.1 mol l-1 dependent on the type of emulsifier. At the CMC most of the emulsifier molecules
have arranged themselves as colloidal clusters with their hydrophobic moiety oriented to the inside
and the hydrophilic moiety oriented to the outside. These so-called micelles can be spherical or rod-
like depending on the type and the concentration of the emulsifier and have a dimension of
approximately 2-10 nm. These micelles will contain the biggest part of the emulsifier concentration,
however there are some molecules dissolved in the aqueous phase [9][2][19][10][26].
When monomer is added to the emulsified solution, a very small portion of the monomer is solved in
the aqueous phase due to limited solubility of the monomer. Another small part of the monomer
molecules, will enter into the micelles, causing them to increase in size. The rest of the monomer
(>95%) will be dispersed in drops into the aqueous continuous phase. The size of these drops is
dependent on the agitation speed and they are stabilized through emulsifier molecules on the drop
surface. The monomer droplets can have a diameter between 1 and 100 µm, which is much larger in
comparison to the monomer containing micelles [9][2][19][26].
To start the polymerization reaction, an initiator is added to the reaction medium. In emulsion
polymerization, the initiator is water soluble in oppose to the initiator used in suspension
polymerization. These water soluble initiators can either be thermal or redox initiators and will
produce free radicals as a result of decomposition in the aqueous phase (see section 1.1.1 above).
Since the initiator is insoluble in the monomer, no polymerization will take place in the monomer
drops. The concentrations of the micelles (1019–1021 L-1) is also much higher in oppose to the
concentration of monomer drops (1012–1014 L-1), giving the micelles the advantage to incorporate
more radicals. Additionally, the micelles possess a much greater totals surface area in comparison to
the monomer drop, making the micelles much more effective in capturing the free radicals.
Polymerization takes place in the aqueous phase. Here, polymerization takes place until they reach
the critical Z-mer length. Subsequently, these short polymer chains will enter in an existing micelle or
will be stabilized by emulsifier in order to form its own micelle. In these micelles, most of the
polymerization takes place. The concentration of monomer in the micelles decreases as the
polymerization proceeds. This decrease is replenished by the monomers in the aqueous phase, of
which the concentration is supplemented by monomer from the dispersed monomer drops
[2][19][26]. A schematic representation of emulsion polymerization is illustrated in Figure 4.

35
Figure 4: General representation of the emulsion polymerization process [2].

In an emulsion polymerization system, three types of particles can be distinguished: monomer drop,
inactive micelles where the polymerization is not (yet) occurring and active micelles where the
polymerization is occurring. The latter particles are often called polymer or latex particles [2].
Formation of the polymer or latex particles can occur according to two different mechanisms. In a
first mechanism, the radicals from the aqueous phase enter the micelle and start the polymerization
reaction. These radicals could either be primary free radicals or propagating radicals (oligomeric
radicals). This mechanism is known as micellar particle nucleation. The second mechanism involves
propagation of monomers in the aqueous phase. These oligomeric radicals become insoluble and
precipitate in the water. The precipitated radicals are now stabilized by the emulsifier molecules,
which were dissolved in the water or adsorbed on the monomer drops, forming an active micelle.
This mechanism is called homogeneous nucleation and is displayed in Figure 5 [2][21].

Figure 5: Representation of homogenous nucleation of active micelles [21].

The extent to which the each mechanism takes place is dependent on the concentration of the
emulsifier and the solubility of the monomer in the aqueous phase. When the concentration of the

36
emulsifier exceeds the CMC, mainly micellar particle nucleation occurs. When the concentration of
emulsifier is approximately the same as the CMC, micellar particle nucleation still dominates
although homogeneous nucleation is also present. However, more homogenous nucleation is
observed as the concentration of emulsifier decreases further below the CMC [2][22].
In general, the emulsion polymerization process is considered to take place in three intervals. The
different intervals can be distinguished by plotting the rate of polymerization and the surface tension
as function of the conversion, as displayed in Figure 6. Figure 7 gives a representation of the
emulsion during the different intervals [2][21][22][26].
In the first interval, particle nucleation occurs and the rate of polymerization increases. The
monomer in the micelles is consumed and replenished from the aqueous phase and the monomer
drops as mentioned above. Due to the formation of polymer in the active micelles, the micelles grow
in size and in number. To stabilize these growing latex particles, they absorb emulsifier molecules
from the aqueous phase. At a certain point, the concentration of emulsifier drops below the critical
concentration to form or maintain micelles (CMC). As a consequence, the inactive micelles become
unstable and disappear by dissolving of emulsifier molecules. Because of the decrease in emulsifier
concentration, the surface tension of the emulsion increases suddenly [2][21][22][26].
In the second interval, the polymerization proceeds in the latex particles with a constant monomer
concentration. This constant monomer concentration is maintained by diffusion of monomer from
the aqueous solution, which in turn is maintained at its saturation concentration by dissolution of
monomer in the water. As a result, the latex particles become larger while the monomer drops
disappear [2][21][22][26].
Because of the absence of monomer reservoirs at the beginning of the third interval, the rate of
polymerization decreases. The rate of polymerization decreases linear because the latex particles still
contain unreacted residual monomers [2][21][22][26].

Figure 6: Surface tension (σ) and rate of reaction (Rp) in function of the conversion for emulsion polymerization [22].

37
Figure 7: Schematic representation of the different intervals during emulsion polymerization [21].

1.2.3 Comparison of suspension polymerization with emulsion polymerization


Suspension and emulsion polymerization have some important advantages over the homogenous
bulk and solution polymerization techniques. Because polymerization processes are exothermic, the
removal of heat from the reactor is important. In emulsion a suspension polymerization, the
monomers are suspended in a continuous aqueous phase, making heat removal much more efficient
in oppose to bulk and solution polymerization. Due to the small drops, a relatively high surface area
to volume is achieved, making the exchange of heat faster. Another advantage of working with
suspension and emulsion polymerization is a lower overall viscosity in comparison to homogenous
polymerization techniques, making agitation and pumping of the reaction medium easier. Thanks to
the easier removal of heat and a lower overall viscosity, high conversions of monomers are possible
whereas these are the limiting factors in bulk polymerization. The conversion of the monomers can
be increased further by the addition of a second initiator or finishing catalyst near the end of the
polymerization [9][2][23][11][28]. In suspension and emulsion polymerization particle size and
particle size distribution can be controlled more easily in oppose to homogenous techniques. This is a
consequence of the choice is suspending agent or emulsifier and variation in agitation rate [9]. These
heterogeneous techniques are often preferred to homogenous technique because these techniques
are able to produce polymers with high molecular weight [2][21]. It is also possible to produce
polymer particles with a core/shell structure in suspension and emulsion polymerization[9][21][11].
This is desired in the production of impact modifiers which consist of a rubbery core and a
thermoplastic matrix. However sometimes the thermoplast and the rubber are not sufficiently
compatible which can be solved by covering the rubbery modifier with a thin layer of polymer which
is compatible with both the rubber and the thermoplast [9].
Despite these advantages of polymerization in a suspension and emulsion, there are some downsides
to this polymerization process. The reaction medium can possess up to 50% water, leading to a
reactor productivity which is much lower in comparison to reactions in bulk or solution
polymerization. This high quantity of water also needs to be separated from the polymer particles in

38
order to obtain the final product. Additionally the present suspending agents in suspension
polymerization and emulsifier in emulsion polymerization need to be removed, because their
presence is often undesirable for certain applications. However, this can make the latex unstable and
lead to coagulation of the particles [2][23][22][24][28].
Although there are many similarities between suspension and emulsion polymerization, there are
some essential differences between the two techniques. First, different initiators are used for
emulsion and suspension processes. In emulsion polymerization, water-soluble initiators are used as
in suspension polymerization mostly water-insoluble initiators are used. This is because initiation
takes place in the aqueous phase in emulsion polymerization and not in the monomer drop as it is
the case in suspension polymerization[9][2][19][26]. There is also a difference in solubility in the
monomers in emulsion and suspension polymerization. In a suspension process, extremely
hydrophobic monomers are desired because the polymerization takes place in the monomer drops.
Loss of the monomer due to dissolving is therefore undesirable because no initiator is present in the
aqueous phase to start the polymerization in oppose to emulsion polymerization[9][2][19][26]. There
is also a difference in rate of polymerization. In emulsion polymerization, the rate of polymerization
is determined by the concentration of the micelles and therefore the emulsifier, whereas in
suspension polymerization the rate of polymerization is dependent on the amount of monomer
drops [9]. Last, the size of the resulting latex or polymer particles is smaller with emulsion
polymerization than with suspension polymerization. These smaller sized polymer particles are a
consequence of the polymerization in the micelles, which slowly swell to latex particles in emulsion
polymerization [9]. In Figure 8 a schematic representation of different polymerization techniques and
their resulting particle size is given.

Figure 8: Schematic overview of different particle sizes in different polymerization techniques [22].

39
2 Polyvinyl chloride
Polyvinyl chloride or PVC is produced from vinyl chloride monomer (VCM). The general reaction is
given by Reaction 17 where n typically varies between 700 and 1500. PVC is one of the most
produced thermoplast in the world because of its inexpensiveness, wide application area and
chemical resistance. Nearly 80% of all PVC produced globally is made by (powder) suspension
polymerization. Also emulsion and bulk polymerization can be used, each contributing 10 % to the
total PVC production [11][26][29][1].

Reaction 17: General reaction of the production of PVC [29].

Two types of PVC are available: flexible or plasticized (pPVC) and rigid (uPVC). They are distinguished
depending on the amount of plasticizer present in the PVC formulation. Flexible PVC is mostly used
for cables, film and floor covering, while rigid PVC is more suitable for window frames, foam and
pipes. Along with plasticizers, other additives are needed to ensure processing and final properties
[29][1][30].

2.1 Additives
Additives are components added to the PVC resins in order to enable processing, improve physical
and mechanical properties, and lower final cost. Basic PVC formulations always contain fillers, to
lower the cost of the PVC composition, heat stabilizers, to improve heat resistance, and colorants,
depending on the desired appearance of the final product [29][1]. In uPVC following components are
often added: processing aids, impact modifiers, and lubricants. In pPVC on the other hand,
plasticizers are the most important additive [1]. The additives mentioned above only form the basis
of standard PVC formulations. Dependent on specific desired properties for PVC, other additives such
as antistatic agents and flame retardants can be added to the PVC matrix[1]. For this work only basic
additives will be relevant and will therefore be explained more in depth in the following sections.

2.1.1 Fillers
The primary reason fillers are added to PVC formulation, is to reduce costs. Through the addition of
the fillers the density of the PVC increases significantly. Because PVC is sold based on weight, the
material cost of PVC will drop with an increasing filler amount [1][31].
Although fillers are added to lower cost in the first place, they can change some properties of PVC. By
adding fillers, properties such as no sticking, increased hardness, reduced tensile strength, reduced
elongation at break, reduction of molding shrinkage, reduction of thermal expansion and abrasion
resistance can be achieved. The addition of fillers can also be accompanied with better processing of
the PVC [29][1][32].
Mainly mineral components such as silicates and silicas, sulphates of the alkaline-earth metals and
calcium carbonate are used as filler for PVC. The latter is definitely the most important and most
used filler. This is mainly due to the wide availability in natural sources, low cost and little energy that
is required for the processing of calcium carbonate. Besides, calcium carbonate has some properties

40
which are desirable for fillers in PVC such as absence of water of crystallization and good resistance
to thermal decomposition during the processing of the PVC composition[1]. Flexible PVC
formulations can contain 30–60 % of filler content. In general, calcium carbonate derived from
natural sources is preferred in flexible PVC formulations due to the less pronounced plasticizing
properties in comparison to precipitated calcium carbonate. In rigid PVC on the other hand, filler
content ranges between 10 to 20 % but in special cases the filler content can rise to 50%. Mostly
precipitated calcium carbonates are used for uPVC and rPVC formulations [1][30].

2.1.2 Colorants
Colorants are mostly added for the esthetic of the final products. They can be divided into dyes,
which are soluble in the PVC formulation, and pigments, which are fine-particle materials insoluble in
the PVC formulation. Dyes are mostly organic compounds and are used to color polymer in
transparent shades. Examples of organic components are (metal-complex) azo-compounds,
anthraquinone, methane, cumarin,… Pigments on the other hand can be divided into two groups,
inorganic and organic components. Inorganic pigments are mainly metal oxides. They are preferred
over other types of colorants because they do not bleed or migrate out of the PVC composition due
to their insolubility. Organic pigments on the other hand do have cleaner and brighter shades in
oppose to inorganic pigments. Some examples of organic pigments are isoindolines, isoindolinones
and dioxazines. In practice, mostly a combination of inorganic and organic pigments is used to obtain
properties of both groups [1][30].

2.1.3 Stabilizers
PVC formulations are prone to degradation under influence of heat or light (UV). Therefore,
stabilizers are added to PVC resin in order to prevent and/or reduce degradation[1]. In the following
sections, the heat and light stabilizers will be explained more in depth.
2.1.3.1 Heat stabilizers
PVC is one of the few thermoplastics which have a higher melting point than its thermal
decomposition temperature. Because of this, heat stabilizers are needed. Under the influence of
heat, HCl is eliminated out of the PVC structure with the formation of polyene sequences, causing
discoloration. This reaction starts at a temperature of 100°C and is represented by Reaction 18
[30][33].

Reaction 18: Representation of the thermal degradation of PVC [1].

The stabilizers exert their function by stabilizing the labile chlorine atom and by neutralizing the HCl
which is released by incorporating a base [3]. Depending on the application for the PVC formulation,
the loading of stabilizer can vary between 0.5 and 8 PHR[1]. There are two types of heat stabilizers. A
first type is the metal-containing stabilizers of which lead stabilizers are frequently used due to their
relatively low cost and longtime stability. However, these lead stabilizers have some limitations. They
are toxic and capable of staining PVC formulations in certain conditions. Tin stabilizers are also widely
used metal-containing stabilizer for PVC. These components are the most effective and long-term
stabilizers for PVC. However, their use is limited due to their high cost and toxicity. A last type of

41
frequently used metal-containing stabilizers is the metallic stearates. These metallic stearates are
mainly cadmium, barium and zinc stearates. Although metal-containing stabilizers are effective,
there are some issues when using these components. They can provoke deposition on metal parts of
the process unit can occur, which is known as plate-out [1][30]. This phenomenon is explained more
in depth in section 3.4.
Also metal-free stabilizers can be used but, these components cannot be used on their own.
However, they can exercise a synergistic effect in combination with metal-containing stabilizers.
metal-free stabilizers are mainly of organic phosphites and epoxy compounds[1][30].

2.1.3.2 Light stabilizers


The degradation of PVC can also occur under the impulse of UV. This process occurs at much slower
rate than it is the case with thermal degradation. Because a large part PVC applications are outdoor
applications, the PVC formulation should be protected against UV or light. The degradation under the
impulse of UV proceeds according to the same mechanism as is displayed in Reaction 18. The
degradation of PVC can also occur via photo-oxidation. This involves the formation of formation of
hydro peroxide, keto, and aldehyde groups via a free-radical mechanism. The formed component are
often also light sensitive and will further breakdown [1][18].
PVC can be protected from UV degradation through the absorption of the UV radiation and re-
emitting this radiation in the visual light spectrum. Titanium dioxide is one of the most used
components for this purpose [1][30][18].
On the other hand, PVC can be protected from UV degradation via the addition of antioxidants.
These components protect the PVC by catching the free radicals and thus preventing the oxidation of
PVC. Components which are used as antioxidant are typically phenol derivatives [1][30].

2.1.4 Plasticizers
More than 90 % of the produced volume of plasticizers is used in the PVC industry. This is due to the
brittle and hard nature of the PVC. Without the addition of plasticizers, PVC would be of little use.
Plasticizers are added to make PVC softer, flexible and better processability. As mentioned before,
PVC can be divided into 2 groups based on the present amount of plasticizer. It is called rigid PVC
(uPVC) when no or little plasticizer is present. When plasticizer is added to the PVC resin, one speaks
of flexible PVC or pPVC [1][3][34].
Plasticizers may reduce the melt viscosity and lower the elastic modulus of the melt. The extent to
which these properties are expressed in the PVC resin is dependent on the type and the
concentration of the plasticizers. The content of plasticizer in pPVC can vary between 20 and 100 PHR
[29][1][34][3].
Plasticizers can be divided into two groups based on their compatibility with the PVC resin. When the
plasticizers are highly compatible with the PVC resin, they are known as primary plasticizers. This
group of plasticizer lowers the glass transition temperature (Tg) and increase the elongation and
softness of the polymer. Secondary plasticizers on the other hand are far less compatible with the
PVC resin and are therefore mostly used in a mixture with primary plasticizers, enhancing the
plasticizing properties of the primary plasticizer [1][34].
The plasticizers can also be classified based on their chemical composition. The largest and most
used group of plasticizers is the phthalates but phosphates are also frequently used as plasticizers. In
addition to the latter two groups, miscellaneous other components such as epoxies and chlorinated
paraffins can be used as a plasticizer [1][34].

42
2.1.5 Polymeric Modifiers
Polymeric modifiers are typically added to the PVC formulation in small amounts. A polymeric
modifier can either be a processing aid or an impact modifier. These components may also have an
influence on the lubricating system. Some impact modifiers for example will increase the
compatibility of external lubricants with the PVC composition. As a result more external lubricant
should be added to the PVC resin in order to become the same lubricating properties as without the
impact modifier [1].
2.1.5.1 Processing aids
In order to promote the processing of the PVC, processing aids are added to the formulation to
improve the melt characteristics. In general, these components melt easier at lower temperatures in
comparison to PVC. As a result, the PVC particles are situated in the viscous melt of the processing
aid, making fusion (see section 2.2.1 in the literature review) more effective and occur much earlier
compared to the none-modified formulations [18].
Mostly acrylic polymers and styrene copolymers are used as processing aids. Most PVC formulations
usually contain just around 1 to 6 PHR of processing aid. Due to this small amount of processing aid,
they will not influence the final properties of the PVC resin drastically [1][18].
2.1.5.2 Impact modifiers
Because of PVC is a brittle material, additives which improve the impact strength and toughness
need to be added to the formulation. These additives are particularly important at low temperatures.
The impact modifiers are dispersed in the PVC structure and are mostly elastomeric copolymers with
low glass transition temperature. The impact modifiers make the PVC tougher through the transfer
of the impact energy to the elastomer. Therefore, a good distribution of the elastomer throughout
the PVC structure is necessary. Also the adhesion at the interface between the PVC and the
elastomer must be strong and preferably by chemical bonding or physical cross-linking. In order for
the impact modifier to be active, they need to be at a concentration between 5 to 15 PHR.
Copolymers which qualify as impact modifiers are for example acrylonitrile/butadiene/styrene
copolymers (ABS) and methacrylate/butadiene/styrene copolymers (MBS) [1][30].

2.1.6 Lubricants
In 2.2, a number of different processing techniques for PVC are described. In order to control
frictional properties and adhesion of the PVC to (metal) parts of the processing unit, lubricants are
added to PVC formulation[3][18].
Because of the importance of these additives for this work, a more detailed description is given in
section 3 in the literature review.

2.2 Processing
PVC can be processed using different techniques such as extrusion, injection molding, calendering
and blow molding in order to obtain the desired shape. The technique which is used is mostly
determined by the final product [1][26][35]. During processing, PVC undergoes gelation or fusion
which is the formation of a homogenous melt out of PVC resin grains. This phenomenon will be
explained more in depth in the following paragraphs. The techniques which are most frequently used
for PVC processing will also be described below.

43
2.2.1 Gelation or fusion of PVC
In order to shape PVC in the desired form, PVC powder or grains need to be melted. This melt is
subsequently transformed into a shape depending on the processing technique. In order to achieve a
homogenous melt from the PVC powder heat, shear and friction are applied to the PVC formulation
[18][36]. A simplistic representation of this gelation or fusion process is given in Figure 9.

Figure 9: Schematic representation of the fusion or gelation behavior of PVC [36].

PVC resin grains or powder grains have a particle size of approximately 100 to 180 µm which are
often referred to as stage III particles. Due the application of heat, friction and shear the PVC resin is
broken down into primary or stage II particles. These primary particles consist out of individual
polymer molecules which form crystalline zones. These crystalline zones are also known as primary
crystallites because they are formed during polymerization. These primary crystallites are connected
by tie molecules as represented in Figure 10 in order to form primary particles [18][36][37][38].

Figure 10: Representation of the PVC structure with A) the crystalline zones connected with tie molecules and B) the
primary particles which are formed out of the crystallites and tie molecules [37].

The polymer molecules of a primary particle can interact with polymer molecules of another primary
particle as a result of their Brownian motion. However, this Brownian motion is rather unimportant
at low processing temperatures because the chains are caught in the crystallites making
interdiffusion and entanglement between molecular chains of two different primary particles almost
impossible. When the temperature rises above 150 °C, due to the application of heat and friction, the
crystallites start to melt. The melting of the PVC is a result of the polymer chains which are able to
escape the crystalline structure and are now free to interact with polymer molecules of other
primary particles as shown in Figure 11A. The primary particles are now able to form a homogenous
melt as their polymer molecules fuse together. When the homogenous melt is cooled, the polymer

44
molecules will form new crystallites which are known as secondary crystallites. These secondary
crystallites ensure the connection between the primary particles resulting in smooth and
homogenous melt [18][36][37][38][37].

Figure 11: Representation of primary particles where A) displays the interaction between primary particles through free
polymer molecules and B) represents the structure after cooling and the formation of secondary crystallites [37].

The gelation or fusion behavior of PVC can be determined by the Brabender gelation test. The
sample in the mixing chamber undergoes fusion or gelation which is similar to the fusion or gelation
taking place during the processing of PVC. In this test PVC grains are masticated in a heated bench
internal mixer. In order to achieve a steady rotation of the rotors to masticate and ‘melt’ the PVC, a
torque is applied. This torque is an indication for the viscosity of the PVC melt and thus the extent to
which the PVC is fused [18]. A more in depth explanation regarding Brabender gelation tests is given
in 2.2.1 in the materials and methods section.

2.2.2 Extrusion
Extrusion is the most used processing technique for PVC formulations. In extrusion, the PVC, in
granular or powder form, is converted into a continuous melt. Subsequently this continuous melt is
forced through a dye with the formation of a product with uniform cross-section. The extrusion
process is typically used for the production of pipes, tubing, film (for packaging), window profiles,…
Due to this broad range of applications, the extrusion process is probably currently the most
important plastic processing technique [1][18].
An extrusion line consists of several different parts. The first and most important part is the extruder.
Here, the PVC granules or powder are melted and forced through a die to achieve the desired shape.
After the extruder, the molten PVC goes through a sizing and cooling unit where the PVC cools down
and the size and the shape of the product are corrected. When the PVC has passed through the sizing
and cooling unit, it reaches the puller which is motor-driven, rubber-covered rolls. The puller ensures
that the PVC is pulled through the sizing and cooling unit to the cutter, which is the last unit of the
extrusion line. In the cutter, the product is cut or sized in the final shape[26]. Figure 12 gives a
general representation of an extrusion line.

45
Figure 12: Schematic representation of an extrusion line with: A) The extruder, B) The die, C) The size and cooling unit, D)
The pull rolls and E) The cutter [26].

As mentioned above, is the extruder the most important part of the extrusion line. In Figure 13 a
representation of the single-screw extruder unit is given. An extruder mainly consists of a hopper
where the PVC granular or powder is brought into the extruder. Via the hopper, the PVC powder is
brought into the barrel, in which a rotating screw is located. By the rotation of this screw, the PVC is
pushed forward in the direction of the die. While the PVC is pushed to the die, it converts from
powder into a homogenous melt and is mixed at the same time. This process is also referred to as
‘gelation’ or ‘plasticization’ of PVC. The heaters located on the barrel, provide sufficient heat in order
to melt and maintain the desired temperature. Also the friction and shear which the PVC is subjected
to contributes to the melting process and maintaining the optimal temperature. The molten PVC is
then forced through the die with the shape of the final product [26][1].

Figure 13: Representation of the extruder [1].

The rotating screw is by far the most important part of the extruder and thus of the whole extrusion
line. Its task consists of pushing the melt forward, mixing this melt and meter the melt to the die. In
general, the rotating screw can be divided into three zones which are distinguished by the difference
in channel depth (see Figure 14 F), which is small at the beginning and increases when closer to the
die. A first zone is the feed zone where the PVC powder is brought into the extruder via the hopper
and is transported into the compression or transition zone. Here, the PVC is compressed in order to
release the present air and plasticized into a continuous melt. This melt is then transported into the
metering zone where the melt is homogenized and a pressure is generated in order to force the melt
through the die at the end of the screw. It should be mentioned that this classification of the zones is
only valid in the case of simple designs of rotating screws such as the one displayed in Figure 14
[26][1][35].

46
Figure 14: More detailed representation of an extrusion screw with: A) The shank, B) The key, D) The helix angle, E) The
root, F) The channel depth, G) The flight, H) The channel, I) The pitch, J) The diameter and K) The tip [26].

2.2.3 Injection molding


Injection molding is generally a widely used technique in the processing of thermoplastics thanks to
its high level of reproducibility and low amount of post-processing steps. In oppose to extrusion, the
final products can have more complex shapes. However, injection molding is much less used in
comparison to extrusion for the processing of PVC. This is due to the thermal instability of PVC and
high melt viscosity. Therefore, the process must be controlled with care [1].
Injection molding is described as a discontinuous process in which the PVC powder is melted and
subsequently injected into a die. Here, the PVC is allowed to solidify through cooling. An injection
molding machine consists of two main parts, the injection unit and the clamping unit, in which the
cyclic steps of the injection molding takes place. In Figure 15 a schematic representation of an
injection molding machine is given. In this figure, the die and the plasticizing unit together form the
injection unit[26][35].

Figure 15: Simplified representation of an injection molding machine with: A) the clamping unit, B) the mold or die C) the
plasticizing unit, D) the control unit and E) the temperature control unit [35].

2.2.4 Calendering
Calendering is a processing technique which is almost exclusively used for PVC. This technique is used
to produce mainly roof-lining material, floor coverings and films from pPVC. The technique is one of
the most expensive ones at this moment due to the cost of the machinery. As a result, calendering is
only economically favorable when producing an excessively high output (40 – 100 m/min or several
tons per hour)[1][35].
Figure 16 displays a calendering line which is commonly used for the processing of plasticized PVC.
The two rolls which are displayed by Figure 16E, are used for the compounding of additives,
plastization and homogenization of the PVC formulation. These two rolls are also known as the roll

47
mills and can also be used as compounding technique for other processing methods. These roll mills
produce a sheet that can be fed in plasticated form directly to the calendering rolls (Figure 16 c) or
after cooling be granulated and fed into an extruder (Figure 16 d) and subsequently to the
calendering rolls [35].

Figure 16:Representation of a calendering line with: a) Winder and edge cutter; b) Cooling rolls; c) Four-roll calender, F
type; d) Extruder; e) Mixing roll mill [35].

Figure 16c represents a four-roll F calender which is mostly used for the processing of pPVC. Using
this conformation, vapor can escape easily and subsequently be removed via suction in order to
prevent condensation on the other roll. Other configurations of the calender rolls are also possible.
Mostly four rolls are used in calendars although they can possess up to seven rolls. These rolls are
commonly made out of highly polished cast iron and are individually driven. In Figure 17 the most
common configurations of calender rolls are given. A distinction is made between I, F, L, or Z
depending on the different setups. In general, F and L type calendars are preferred for the processing
of PVC because they are fairly cheap and easy to use. The temperature and the speed in the calendar
are generally increased from one roll to the other. The temperature is controlled through the
individually heating of the rolls with the use of oil or water as a heating medium [1][35].

Figure 17: Different configurations of calender rolls with A) I-type; B) F-type; C) L-type and D) Z-type calendars [35].

The PVC is fed continuously from the extruder to the first slit of the calendering rolls. As a
consequence of the pressure in this slit, a built up of the PVC is created. The sheet on the other hand
can pass through the slit unaffected. As a result, the melt is more homogenized through the
influence of the created currents (Figure 18). The quality of the product is therefore determined by
the formed built up or kneading stock. The built up near the first slit can be influenced by the
adjustment of the individual rolls [1][35]. The cross section of this built up and the currents which are
created in by this built up are displayed by Figure 18.

Figure 18: Representation of the cross section of the built up and created currents with: A) Kneading current; B) Feeding

48
current; C) Outlet current [35].

When the sheet has left the calender, it is arrives in the cooling unit (Figure 16b) where the sheet is
cooled down until it is stable and does not deform anymore. Finally, the sheet reaches the winder
and edge cutter where it is cut into the desired shape and/or rolled up (Figure 16a). Films or sheet
which are produced through calendering are typically 1.5 m wide and have a thickness between 0.1
and 1 mm[35].

2.2.5 Blow molding


Blow molding is a processing technique used to produce hollow objects out of PVC. There is a variety
of products that can be produced via injection molding, but the most known is the production of
bottles[26][1][35].
In blow molding a parison, which is basically a hollow tube from a molten PVC resin, is enclosed
between a die consisting of two parts. Subsequently, with the use of air under pressure (6.89 bar),
the parison expands against the cold walls of the mold. The object is cooled down and is removed
from the mold. Afterwards, the excess PVC can be removed to obtain the desired shape[26][35].
Depending on the way the parison is created, two different processes of blow molding are
distinguished. The first process is known as extrusion blow molding. In this process, an extrusion
process (as described in2.2.2) is used to make the parison. Figure 19 represents the consecutive
actions which are executed in extrusion blow molding. When the extruded parison has reached the
desired length (Figure 19A), it is placed in the mold by a clamping mechanism, the mold is closed
(Figure 19B) and then transported to the blowing station while the parison is still hot (Figure 19C).
Here, the air under pressure is injected into the mold in order to expand the parison to its final form
(Figure 19D). When the object has cooled down, it is removed from the mold (Figure 19E)[26][35].

Figure 19: Representation of extrusion blow molding line where: A) the parison extrusion; B) positioning of the blow mold;
C) positioning in the blow station; D) Shaping and cooling; E) Removing of the object from the mold [35].

When the parison is created by injection molding, the process is called stretch blow molding or
injection blow molding. In this process, the injection molded parison is placed in the mold and the
process proceeds the same way as for the extrusion blow molding [26][35].

49
3 Lubricants
Lubricants are additives for polymers which improve the processability. The processability is
especially a problem with PVC formulations because of its high melt viscosity. As a result of this high
melt viscosity, high power and heat consumption is needed in order to process PVC. Besides, the high
melt viscosity causes the PVC to stick to metal parts of the processing unit. To avoid and reduce
these problems, lubricants are added to the PVC formulation. Lubricants will reduce energy
consumption, melt pressure in the process unit and improve dispersion of other additives such as
fillers and pigments. They can express their function through two different mechanisms: internal and
external lubrication. Most lubricants are however able to express both internal and external
lubrication [1][30][3][4][5]. Internal and external lubrication will be discussed more in depth in 3.1
and 3.2.
Lubricants can lower the friction due to the presence of their long hydrocarbon chains (>C12). Their
properties are dependent on the solubility in the PVC matrix and the interaction between the metal
surface and the lubricant [1][30][33][4][39].
Depending on the type of PVC, a different lubrication system is needed. In flexible PVC for example,
only external lubricants are needed, whereas the present plasticizers in the flexible PVC have enough
internal lubricating power to ensure good processing. For the processing of rigid PVC on the other
hand, external as well as internal lubricants are always needed [1][30][39].
The amount and the type of lubricants are strongly dependent on the processing and the used heat
stabilizer. When for example lead stabilizer is used, only a small amount of lubricant is sufficient due
to the strong internal lubricating properties of the lead stabilizers. Tin stabilizers on the other hand,
require more lubricant because they tend to stick to the metal surfaces of the machinery. In general,
the processing of PVC with injection molding also requires more additives in comparison to extrusion
[39].

3.1 Internal lubricants


Internal lubricants reduce intermolecular friction during the gelation of the PVC. Due to this
reduction of friction, internal lubricants reduce the melt viscosity and therefore improve the flow
properties of the PVC formulation. As a result, the output of the processing can significantly increase
and the processing is subjected to less strict conditions [1][30][39].
In general, internal lubricants have a high compatibility with the PVC matrix, due to their more polar
structure. These lubricants perform their function by blending into the body of the PVC through
dissolving in or bonding strongly to the surface of the polymer molecules. Thus, the internal lubricant
forms a continuous layer of molecules around the PVC molecules, which eases the movement
between two molecules or polymer particles as well as the movement of the molecules or polymer
particles with the metal surface of the processing unit. This results in a reduction of the ‘internal’
friction and has proven to have a positive influence on the viscosity of the melt, the softening
temperature of the PVC and buildup of heat during processing[1][30][4][5][39]. In Figure 20 the
internal lubrication of PVC is represented by the red stripes.

50
Figure 20: Representation of internal and external lubrication in PVC processing [3].

3.2 External lubricants


External lubricants are added to PVC formulation in order to reduce adhesion and friction to the hot
metal surface of the machinery. As a consequence the abrasion between the metal surface and the
PVC melt is reduced and thus the melt flow of the PVC is improved. Good external lubrication hence
ensures that the PVC melt can glide easily over the machinery parts.[1][30][4][5][39].
In oppose to internal lubricants, external lubricants mostly have low compatibility with the PVC
matrix. Due to its incompatibility, external lubricants can go out of the PVC matrix to the metal
surface of the processing unit. The extent to which the external lubricant can go outside, and thus
exert its external lubricating properties, is dependent on the degree of incompatibility. Due to the
external lubricant, an increase of the melt flow velocity at the metal surface of the machinery is
observed (Figure 21b). This is not the case with internal lubrication which results only in a higher flow
of the polymer melt front in the middle of the die or barrel (Figure 21a). [1][30][5][39].

Figure 21: The influence of A) internal and B) external lubrication on the speed of the flow front of the polymer melt [39].

An important factor when using external lubricants is the amount at which they are added to the PVC
matrix. When too much of the external lubricant is added, overlubrication can occur which causes
slippage. This is not desirable because there is minimum of friction on the surface of the metal
needed in order for the PVC to move smoothly through the barrel of the machinery. For this reason
the concentration of external lubricants must be carefully controlled [30][5].
Figure 20 schematically displays how external lubricants express their function. The external
lubricant is represented by the green lines.

51
3.3 Current lubricants
Lubricants are classified in five groups based on their chemical nature because classification based on
internal and external behavior is not clear since most lubricants are able to exert both [33].
A first class of lubricants are the amides. In particular, ethylenebisstearamide (EBS) is used for the
lubrication of rPVC. EBS is also known as amide wax and is derived from stearic acid, origination form
animal or vegetable fats or oils and ethylene diamine. EBS has good internal as well as external
lubricating properties. However its use has decreased over the years to the change from single to
twin screw extruders which require more external lubrication than provided by EBS[33].
A second group which is distinguished are the hydrocarbons or the so called synthetic waxes. These
components are produced by the polymerization of small hydrocarbon monomers such as ethylene
and propylene As a result, this group mainly consists of polyethylene, polypropylene and paraffin
waxes. These lubricants all have the following general structure: -(CH2)n- where n can vary from 20 to
80. This group of lubricants mainly express external lubricating properties and do have good metal
release characteristics. An additional internal lubricant for the processing of rPVC is thus required
when using these hydrocarbon waxes [30][33][3][39].
Also esters can be used as lubricants in the processing of rPVC. This group consists of a variety of
different esters such as glycerol ester and montan esters. These esters are derived from alcohols and
fatty acids. Due to the large variety of esters, they express internal as well as external lubrication
behavior. In general the external lubricating properties of these esters increase with an increasing
chain length [33][3] [39].
Fatty acids are also available as lubricants. The most known fatty acid used for lubrication of rPVC is
stearic acid. This component has good external lubrication and some internal lubrication. However,
due to its volatility its use is very limited[33][3][39].
Last, metallic soaps are used for their excellent lubricating properties. These components are mostly
soaps of alkaline earth metals. Calcium stearate is the most used metallic soap as a lubricant. It is
known to improve flow and mold release and thus provides internal and external lubrication.
However, it has been demonstrated that it can increase internal shear which is not desired from the
perspective of internal lubrication. Therefore, calcium stearate is mainly used as a colubricant instead
of on its own. It is almost used in every PVC formulation. However, the use of calcium stearate has a
big disadvantage. It is known to increase plate-out together with hydrocarbon waxes and heat
stabilizers during the processing of PVC[33][3][39]. In 3.4 a more detailed description of plate-out is
given.
In general, none of the classes above are used on its own because of the complexity of the
processing of rPVC. The current processing is in need of a balanced lubrication system which can
provide the PVC with good internal and external lubricating properties. Therefore a mixture of two or
more of the above mentioned lubricants are mostly used[33][39].

3.4 Plate–out
Plate-out is one of the most distinct problems during the processing of PVC. Plate-out is the
uncontrolled deposition on metal parts of the processing unit under the influence of lubricants and
stabilizers. Mainly hydrocarbons lubricants are known to cause plate-out [7][40][6]. Especially during
the extrusion of PVC plate-out seems to be a problem. Deposition in extrusion units mainly occurs in
places where changes is pressure and flow occur, such as on the wall. As a result of these deposits,
there will be marks on the final extrusion product, which is obviously undesired. As a consequence,

52
plate-out causes material wastage, and loss of time and output [1][41].
The exact knowledge of how plate-out arises is not fully clear. However, it has been discovered that
formulation as well as process conditions have major influence on the formation of plate-out. The
research of Lippoldt has shown that plate-out consists of inorganic, which were mainly metals from
stabilizers, as well as organic compounds, which mostly originated from the lubricants. This led to
believe that plate-out is a consequence of a poor ratio between stabilizers and lubricants. Lippoldt
experimentally established a mechanism which is most likely to cause the formation of plate-out.
Lippoldt claims that the lubricants, which are mostly hydrocarbons, dissolve inorganic additives and
stabilizers in the PVC melt. When the temperature in the processing unit reaches about 175 °C the
colubricant calcium stearate dissolves in the hydrocarbon-stabilizer mixture and forms a complex.
This complex tends to adsorb on the surface of the inorganic additives, making them less polar.
When this complex of inorganic stabilizer, hydrocarbons and calcium stearate reaches a
decompression zone in the processing unit, the temperature drops below 175°C resulting in the
precipitation and deposition of this complex. These depositions now form a site at which material
deposit can continue [1][7][40][6].
Plate-out is strongly affected by the choice of lubricant and stabilizer, in combination with the
applied processing conditions. It is for example recommended to use a lubricant which does not
dissolve the stabilizer[1].

3.5 Acrylic Lubricants


As mentioned above, the properties of lubricants are determined by their compatibility with the PVC
matrix, which is determined by the polarity of the component. The more the lubricant is
incompatible, and thus non-polar, with the PVC matrix the better the external lubricating properties.
This is a consequence of a weak bond strength between the lubricant and the PVC, making it possible
for the lubricant to escape the PVC matrix under the influence of high shear. The external lubricant
then lubricates the PVC by covering the metal surfaces of the processing unit[1][33][5][39]. However,
this incompatibility of the lubricant with the PVC matrix can contribute to the formation of plate-out
as it is the case with hydrocarbon waxes such as polyethylene and propylene waxes [7][40][6].
Therefore, the use of internal lubricants is preferred because of their compatibility with the PVC
matrix. They will arrange themselves under the influence of shear in the direction of the flow,
reducing shear stress between the PVC molecules. This will cause much less plate-out in oppose to
classic external lubricants because they will not escape the PVC matrix [33].
Because for the processing of rigid PVC both internal and external lubrication are required, the use of
solely an internal lubricant is not a solution [33]. For this reason, there is need for an external
lubricant which does not completely migrate out of the PVC structure. This type of lubricant is, in
accordance with acrylic processing aids, a low molecular weight copolymers of various methacrylate
monomers [42][43].
There are some synthetic acrylic lubricants available on the market but they are mostly oil-based and
used in different application than envisaged in this study [43][44]. As a result no mechanisms of how
acrylic lubricants work in PVC processing have been found.
Because the acrylic lubricants has a relatively polar structure thanks to the presence of the ester
function, it is believed that these components will have a high compatibility with the polar PVC
matrix [45][46]. This compatibility with the PVC matrix is even improved due to their low molecular
weight and thus short chains. The short chains of the acrylic lubricant lead to a lower viscosity,
making it easier for the acrylic lubricant to disentangle from the PVC matrix. As a result, the acrylic

53
lubricant can go easily to the outside of the PVC matrix. However, due to the good compatibility with
the PVC matrix, it will not separate completely and form a second layer which is the case with the
current acrylic lubricants [39][47][48].
Because of the latter two properties of the acrylic lubricant in the PVC matrix, it is believed that the
problem with plate-out will be reduced significantly. The acrylic lubricant will execute external
lubrication by going to the outside under influence of shear in the processing unit. Here, it will
reduce the friction with the metal surface of the processing unit but will not completely separate
from the PVC matrix [39][47][48].

54
PART 2: MATERIALS AND METHODS

1 Materials
1.1 Materials for polymerization
Reagents used during the polymerization with their function are given in Table 1. For confidentiality
reasons, no structures or properties are given. All materials were obtained from Arkema or Sigma-
Aldrich.

Table 1: Chemicals used during polymerizations with their function

Component Function
Monomer M Monomer
Monomer L Monomer
P7 Initiator
P4 Initiator
A81 Initiator
E/H6 Reducing agent
F Recovery reducing agent
H16 Buffering agent
EM15 Emulsifier
EM17 Emulsifier
M1 Chain transfer agent

The reagents with their function, in order to perform the coagulation (see section 2.1.3 in materials
and methods) are given in Table 2. For confidentiality reasons, no structures or properties are given.
These materials were also derived from Arkema or Sigma-Aldrich.

Table 2: Chemicals with their function which are used during coagulation of the latex

Component Function

H8 Electrolyte to destabilize latex


EM12 Emulsifier to stabilize the formed particles
STA12 Emulsifier to stabilize the formed particles

1.2 Materials for evaluation


For the determination of the lubricating properties of the acrylic polymer, a standard PVC
formulation was used: NWG. This formulation consists of PVC, an internal lubricant, an external
lubricant and stabilizers. A more detailed description of this formulation is given in 2.1 in the results
and discussion section.

55
2 Methods
2.1 General set-up and procedure of the polymerization experiment
The aim of this study is to develop a low molecular weight acrylic polymer which exerts external
lubricating properties. As mentioned in the literature review (section 3.5), the acrylic lubricant needs
to have a low molecular weight in order to express its external lubricating properties in PVC
formulations. This acrylic polymer was synthesized via free radical suspension polymerization. In
order to achieve this low molecular weight, different parameters of the suspension polymerization
process were changed. Parameters such as temperature, initiator type, initiator concentration,
emulsifier type, emulsifier concentration and addition of a chain transfer agent were adapted. For
this research, reaching the target molecular weight is the main goal. Nevertheless, the influence of
the parameters on the total monomer conversion, the conversion rate, the solid content of the latex
and the particle size (distribution) are also important with respect to the polymerization as well as
the final properties of the polymer. A general description of a recipe with the added components and
their function is given in section 2.1.2 in this section. An overview of the adaptations made to this
recipe is given in Table 37 in the attachment.

2.1.1 General set-up


All polymerization experiments were carried out using the set-up displayed in Figure 22. For every
experiment, two (semi-) reactors were used in parallel with a slight variation in recipe, indicated with
‘1’ and ‘2’. However, the set-up of both reactors is the same.
The main part of this set-up is a 5 liter reactor (Figure 22.3), which is placed into a water bath in
order to control the reaction temperature (Figure 22.1). This bath is heated using two Julabo ED
heaters (Figure 22.2). During the polymerization reaction, the reactor is stirred at 250 rpm via a
Teflon propeller stirrer with a diameter of 8.0 cm (Figure 22.4). This stirrer is placed at a height of
approximately 1.5 cm from the bottom of the reactor and is driven by an IKA-WERKE EUROSTAR
digital overhead stirrer (Figure 22.5). In order to optimize mixing, two finger baffles (Figure 22.6) are
placed into the reactor to break the flow. Temperatures of the bath and reactor are measured during
polymerization using thermocouples (Figure 22.7). These temperatures are recorded by a Yokogawa
view recorder (Figure 22.8). In order to avoid evaporation of monomers, water,… a reflux cooler
(Figure 22.9) supplied with cooling water is attached to the reactor.
In order to prepare the monomer emulsion with a volume average drop size of ± 1 µm, a Tokushu
Kika TK Homo mixer is set at a rotational speed of 10 000 rpm for 5 minutes. During polymerization,
this monomer emulsion is transported from the monomer vessel (Figure 22.10) to the reactor using a
FMI Lab pump (Model QG 105) (Figure 22.13) at the desired fixed dosing rate. The monomer
emulsion is continuously stirred during the polymerization by a 2 Mag magnetic motion MIX drive
magnetic stirrer (Figure 22.11) at 400 rpm, in order to avoid phase separation. The monomer vessel,
as well as the reactor, is kept under a nitrogen flow of approximately 100 ml/min (Figure 22.12) to
exclude oxygen as much as possible from the polymerization reaction.
For some experiments, an additional FMI Lab pump (Model QG 105) (Figure 22.15) is installed in
order to obtain a continuous dosing of a second component (Figure 22.14) during the polymerization.
This vessel is not kept under a nitrogen atmosphere because this is not feasible from the industrial
point of view.

56
Figure 22: General set-up of the polymerization reactions with 1) the water bath, 2) the heaters, 3) the reactor, 4) the
stirrer, 5) the stirrer driver, 6) the baffles, 7) a thermocouple, 8) the temperature recorder, 9) the reflux cooler, 10) the
monomer vessel, 11) the magnetic stirrer, 12) the nitrogen supply, 13) the pump for the monomer emulsion, 14) the vessel
for the second component and 15) the pump for the second component.

2.1.2 Polymerization experiment


The polymerization recipe exists of a series of charges which are added to the reactor. The
composition of these charges and their corresponding concentrations are given in Table 3. All
concentrations are given in PHR or parts per hundred rubber or resin. This is the amount of
compound that is added per hundred parts of the polymer. For confidentiality reasons, some
concentration may be encoded with letters. All amounts are then expressed relative to the final
amount. X indicates however a variable concentration in Table 3.

Table 3: Overview of the components with their function and concentration (PHR) for every charge added during the
polymerization experiment with X representing a variable concentration

Concentration
Charge Components Function Time of addition
(PHR)
Continuous
PW 52.6
phase
Charge 1 (= initial Start of the
Regulation of
charge) H16 0.05 polymerization
pH
(EMx) Emulsifier X
Reducing
E/H6 0.01
agent
5 minutes after
Charge 2 (= Recovery
reaching reactor
initiation charge) F reducing 0.1
temperature
agent
P7 Initiator R
Continuous
PW 100.0
phase
M Monomer A 5 minutes after
Charge 3a (=
L Monomer B charge 2
monomer
EMx Emulsifier X Polymerization time =
emulsion)
Chain transfer 0 minutes
M1 X
agent
P4/A81 Initiators X

57
Concentration
Charge Components Function Time of addition
(PHR)
5 minutes after
Recovery
charge 2
Charge 3b F reducing X
Polymerization time =
agent
0 minutes
30 minutes after start
Charge 4 P7 Initiator 0.14 R
of monomer addition
Approximately 110
End of Monomer
minutes after charge
dosing
3a
Regulation of 120 minutes after
Charge 5 H16 0.1
pH charge 4
End of 60 minutes after
polymerization charge 5

The charges given in Table 3 are added into the reactor using the following method: the
polymerization is started by loading the initial charge (charge 1) into the reactor which is afterwards
heated to desired temperature using an isothermal water bath. The reactor with the initial charge
and the monomer emulsion (with initiators) are vented with nitrogen gas at a rate of 600 ml/min.
Afterwards, the flow rate of the nitrogen is reduced to 100 ml/min. Five minutes after reaching the
desired reactor temperature the initiation charge (charge 2) is added into the reactor. In order to
avoid penetration of oxygen into the reactor, the nitrogen flow on the reactor was raised to 300
ml/min. Due to the addition of the reagents, which are at room temperature, a significant drop in
reactor temperature is noticed. In order to ensure the formation of sufficient radicals, monomer
addition (charge 3a) started five minutes later. The moment the monomer emulsion reaches the
reactor is set as time zero for the polymerization. Simultaneously with the start of the monomer
addition, the addition of a second component can be started. The rate at which these components
are added to the reactor can be different dependent on the recipe. Thirty minutes after starting the
addition of the monomers (and the second component) charge 4 is added to the reactor. The
monomer dosage ends approximately 110 minutes after the start of the addition. Subsequently,
charge 5 is added 120 minutes after charge 4. Another 60 minutes later, the polymerization is
complete.
In order to evaluate and check the progression of the polymerization, samples are taken at various
times. Table 4 displays at which time samples are taken accompanied with the tests performed on
the samples. During sampling, the nitrogen flow is increased to 300 mL/min in order to avoid the
introduction of oxygen into the reactor. The different tests which are indicated in Table 4 are
explained more in depth in section 2.1.4 below.

Table 4: Overview of the time (min) and tests performed on the samples during polymerization, with SC the solid content,
ƞsp the specific viscosity and GC gas chromatography

Sample Polymerization time (min) Test


Sample 1 25 SC, ƞsp, GC
Sample 2 60 SC, ƞsp, GC
Sample 3 90 SC, ƞsp, GC

58
Sample Polymerization time (min) Test
Sample 3’ Just after monomer dosing has SC, GC
ended
Sample 3” 30 min after S3’ SC
Sample 4 (final) 180 SC, ƞsp, GC

After polymerization, the obtained latex is poured through a sieve (35 mesh) in order to check the
formation of scales, which is the undesired and uncontrolled coagulation of latex particles, during the
polymerization.

2.1.3 After treatment


In order to achieve powder out of the final latex, a controlled coagulation process is used. 333.3 PHR
latex is diluted with 333.3 PHR purified water. Subsequently, V PHR H8 in 333.3 PHR purified water is
added while stirring the mixture on a Labinco L32 stirring and heating plate. In order to stabilize the
formed particles, Z PHR EM12 is added to the mixture while stirring is maintained. Then, the mixture
is heated to T °C and afterwards filtered with a 1 l Buchner flask, a KNF Lab Laboport pump and
Whatman 589/1 filter with a pore diameter of 185 mm. The obtained powder is dried overnight in a
Heraeus vacuum oven at room temperature. Afterwards, the powder is placed in the freezer and
subsequently crashed into a fine powder using a mortar.

2.1.4 Analysis of the samples


2.1.4.1 Molecular weight determination
A qualitative idea of the molecular weight of the acrylic lubricant can be obtained by measuring the
specific viscosity (ƞsp). This is done by dissolving the polymer into an appropriate solvent. As a result,
the mobility of the solvent molecules is reduced and the viscosity of the solution is increased. The
specific viscosity can be determined according to the following equation:
𝜂𝑠𝑜𝑙𝑢𝑡𝑖𝑜𝑛 − 𝜂𝑠𝑜𝑙𝑣𝑒𝑛𝑡
𝜂𝑠𝑝 =
𝜂𝑠𝑜𝑙𝑣𝑒𝑛𝑡
In this equation, ηsp the specific viscosity, ηsolution the viscosity of the solution and ηsolvent the viscosity
of the solvent [9].
To determine the specific viscosity, measurements were conducted with a Viscotek model 340 which
is an automatic system capillary type viscometer, in combination with ARV software (version 2.0.7).
The samples were prepared by drying the latex at a temperature of 100 °C for one hour on a petri
dish or using the coagulation process described above (2.1.3 in the materials and methods section).
Afterwards, a solution of exact 0.3984 g/dl (0.4 %) was made in toluene and is stirred for 90 minutes
at a temperature of 30°C. Afterwards, 25 ml of the solution is injected via auto sampling and the
specific viscosity is measured.
The results of the specific viscosity measurements were set in relation to gel permeation
chromatography (GPC) or size exclusion chromatography (SEC) measurements. This is a well-known
technique for the determination of the molecular weight (distribution) of polymers.
To measure GPC 2 to 3 mg of the latex residue (as described above) is added in 1 ml tetrahydrofuran
(THF) and stirred overnight to ensure the sample to be completely dissolved. 100.00 µl from this
solution was injected in a TOSOH H7000 column. The temperature was kept a 40 °C throughout the
column. The flow was set on 0.45 mL/min resulting in a pressure between 62.05 bar and 68.95 bar.
The molecules which are eluted from the column are detected through a differential refractometer,

59
viscometer and light scattering detector which measure the amount of polymers has passed through
the column as function of time. Out of the acquired data, a molecular weight distribution is
determined.
2.1.4.2 Solid content
In order to determine the solid content (SC), a CEM Square sample pad is weighed on a Sartorius
LA3200D scale. Subsequently, 3 mL of the latex was added onto the sample pad. The sample pad is
then placed into a Whirpool Compact microwave for 5 minutes at 500 W in order to remove the
water. Afterwards, the sample pad is weighed and the solid content is determined according to the
following equation:
𝑤𝑒𝑖𝑔ℎ𝑡 𝑑𝑟𝑦 𝑠𝑎𝑚𝑝𝑙𝑒 𝑝𝑎𝑑 − 𝑤𝑒𝑖𝑔ℎ𝑡 𝑤𝑒𝑡 𝑠𝑎𝑚𝑝𝑙𝑒 𝑝𝑎𝑑
𝑆𝐶% = × 100
𝑤𝑒𝑖𝑔ℎ𝑡 𝑤𝑒𝑡 𝑠𝑎𝑚𝑝𝑙𝑒 𝑝𝑎𝑑 − 𝑤𝑒𝑖𝑔ℎ𝑡 𝑠𝑎𝑚𝑝𝑙𝑒 𝑝𝑎𝑑
2.1.4.3 Residual monomer
To determine the residual monomer in the latex, gas chromatography is used. Samples of the latex
are prepared in 4 mL vials. The vial is filled with approximately 2 mL of diethyl ether by Sigma-
Aldrich. Subsequently, 50 µL of an internal standard is added, 2880 µg/ml chlorobenzene in diethyl
ether. A known amount of latex (approximately 0.500 g) is then added. The vial is shook on the
Specamill for 30 minutes. Afterwards, the vials are placed in a Hettich® EBA 21 centrifuge for 15
minutes at 5000 rpm.
After centrifuging, a phase separation has taken place between the solvent and the latex and the
residual monomer is located in the solvent. Subsequently, exact 1 µL of the solvent layer, is injected
on column using a split-splitless injector in a Varian CP-3800 gas chromatograph which is driven by
CompassCDS software. This gas chromatograph is equipped with a flame ionization detector (FID).
The temperature program and column used to determine the residual monomer is dependent on the
monomer.
For the determination of the residual monomer L, a HP-FFAP from Agilent with a length of 15 m, an
internal diameter of 0.53 mm and a film of 1.0 µm. The temperature program which is used in order
to determine the residual L is given in Table 5. The pressure is kept constant at 0.21 bar.

Table 5: Temperature program (°C) for the determination of residual L on a GC with a HP-FFAP column

Unit of the GC Time (min) Heating rate (°C/min) End temperature (°C)
Injector 0-1 0 40
Injector 1 – 8.5 25 230
Injector 8.5 – 60 0 230
Column oven 0-1 0 40
Column oven 1 - 20 10 230
Column oven 20 - 60 0 230
Detector (FID) 0 - 60 0 250

For the determination of the residual monomer M an AT-WAX COLUMN from Alltech Heliflex with a
length of 30 m, an internal diameter of 0.32 mm and a film of 1.0 µm is used. The used temperature
program is given in Table 6. The pressure is kept constant at 0.62 bar during the analysis.

60
Table 6: Temperature program (°C) for the determination of residual M on GC with an AT-WAX column

Unit of the GC Time (min) Heating rate (°C/min) End temperature (°C)
Injector 0-42 0 150
Column oven 0-1 0 50
Column oven 1 - 11 5 100
Column oven 11 - 17 0 100
Column oven 17 - 27 10 200
Column oven 27-42 0 200
Detector (FID) 0 - 42 0 250

Out of the peak areas for the internal standard and the monomers, the residual monomer in the
latex can be determined. In Table 7 and Table 8 the retention times of the monomers and the
internal standard are given for both columns using the temperature profile as mentioned above.

Table 7: Retention times (min) of L and the internal standard using a HP-FFAP column and the temperature profiles
described in Table 5

Retention time
First retention time L Second retention time Third retention time L
internal standard
(min) L (min) (min)
(min)
13.5 15.4 17.3 3.5

Table 8: Retention time (min) of M and the internal standard using an AT-WAX COLUMN and the temperature profiles
described in Table 6

Retention time M (min) Retention time IS (min)


6.20 12.16

Using the following equation the total amount of residual monomer can be determined:
𝐴𝑀𝑀𝐴 ∑ 𝐴𝐿𝑀𝐴
𝐴𝐼𝑆 × 𝑓𝑀𝑀𝐴 × 𝐶𝐼𝑆 + 𝐴𝐼𝑆 × 𝑓𝐿𝑀𝐴 × 𝐶𝐼𝑆
𝑚𝑠𝑎𝑚𝑝𝑙𝑒 × 𝑆𝐶
𝑚=
∑ 𝐴𝐿𝑀𝐴
𝐴𝐼𝑆 × 𝑓𝐿𝑀𝐴 × 𝐶𝐼𝑆
1− 𝑚𝑠𝑎𝑚𝑝𝑙𝑒 × 𝑆𝐶
In this equation m represents the mass of the residual monomer in µg monomer per gram polymer,
AM and AL represent the peak area for respectively monomers M and L, AIS represent the peak area
corresponding with the internal standard. fM and fL are the response factors, CIS the concentration of
the internal standard in µg/ml and SC the solid content. The factor f is dependent on the response of
the detector on the monomer, and is determined 1.400 for M and 0.972 for L.
Using the residual monomer content, the actual monomer conversion and rate can be determined
when the monomer dosing time is known.
2.1.4.4 Particle size distribution
The particle size was determined because of its influence on the lubricating properties and to get an
idea of the reproducibility of the reaction and how the reaction went. Particles which have a particle

61
size between 0.1 and 10 000 µm were determined with the Microtrac S3500 with Microtrac FLEX
10.5.4 software using the mode for absorbing spherical particles. The Microtrac S3500 uses laser
diffraction to determine the particle size volume distribution. With the Microtrac S3500, it is possible
to measure latex as well as powder. When particle size distribution of latex is measured, one or two
drops of the latex were added to the reservoir, depending on the loading factor which is displayed by
an indicator bar. When measuring powder, two grams of the powder needs to be diluted with 100
mL of a 2 % solution of EXTRAN MA01 to ensure complete solution of the powder particles. This
mixture is then stirred for 2 minutes in order to homogenize. Subsequently, this dilution is added to
the reservoir until proper loading is reached. Both samples are measured for 60 seconds.
For particles which have a size in the range of 0.0001 µm and 10 µm the Nanotrac Wave ll was used
with Microtrac FLEX 11.1.0.3 software in transparency mode. This instrument uses diffractive light
scattering to determine the particle size intensity distribution. For these measurements, a 0.2 %
solution of the polymer in water is made from the latex. Then the sample cell is filled with the
solution and the sample was measured three times for 240s.
Out of the measurements with the Microtrac S3500 and the Nanotrac Wave ll, the D 50 value can be
determined. This means that 50 % of the particles are smaller and 50 % of the particles are larger
than this D50 value. To confirm the measurements of the Microtrac S3500 and the Nanotrac Wave ll,
the particles size was determined using scanning electron microscopy.
SEM pictures were taken by adding a small amount of the coagulated powder (see section 2.1.3 in
materials and methods) on a sample holder which contains conductive tape. Onto this sample a layer
of gold is deposited, which distributes the electrons over the complete sample. Subsequently, this
sample is placed into the jeol JSM-6010LA scanning electron microscope with which images can be
generated in different magnifications. Using the software a number particle size distribution can be
obtained.
2.1.4.5 Glass transition temperature
The glass transition temperature was determined using differential scanning calorimetry (DSC).
Based on the glass transition temperature, it is possible to clarify the formation of a homopolymer or
a copolymer.
For the DSC measurements, 5 – 10 mg polymer powder is added into a 40 µl aluminum pan and
placed into the Mettler Toledo Sample Robot TSO801RO. This auto sampler places the pan into the
Mettler Toledo DSC 822e. A DSC curve is obtained via the Mettler Toledo STARe V13.00 software. The
temperature profile which was used is given in Table 9.

Table 9: Temperature profile (°C) used during DSC measurements

Temperature (°C) Time (min) Heating rate (°C /min)


- 60 5 /
140 20 10
-60 20 -10
140 20 10

62
2.2 Lubricating properties
The final lubricating properties of the acrylic polymers are determined using Brabender gelation and
capillary rheometry.

2.2.1 Brabender gelation test


2.2.1.1 General set-up
In order to define the influence of the synthesized acrylic polymers, a Brabender Plasticorder PL 2000
was used to determine the gelation behavior of a PVC formulation with and without the synthesized
polymers. The gelation behavior of PVC was already discussed in section 2.2.1 part. The Brabender
Plasticorder was used in combination with roller mixing blades and the Brabender Mixer Program
(Fusion behavior) V4.9 software. The general set-up and the different components of the machinery
are given in Figure 23.

Figure 23: General set-up for the Brabender gelation tests with A) the completely assembled equipment and B) the
different components of the disassembled chamber[36].

2.2.1.2 Procedure
There are three parameters which can be adjusted in order to influence the gelation behavior of the
PVC. A first parameter is the temperature of the chamber. By increasing the temperature, gelation of
the PVC will occur sooner. A second parameter is the rotational speed of the mixing blades. By
increasing the rotational speed, friction and shear on the PVC will increase causing gelation to start
earlier. The last parameter that influences the gelation behavior is the amount of powder added to
the chamber. When increasing the chamber filling, more shear and friction is applied to the PVC
causing it to gelate sooner [18].
For the conducted experiments, the temperature of the chamber was set to 160 °C and the rotating
speed of the roller mixing blades was set at 30 rpm. The formulation was weighed on a Sartorius
Laboratory LC420 balance and acrylic polymer was added cold to this formulation. Subsequently, this
mixture was mixed for one minute using an IKA-WERK TYPRW 18 mixer. The mixing chamber was
filled with 50.0 or 54.0 grams of a PVC formulation, depending on the experiment, via the hopper
and using a plunger or pestle. When the mixing chamber is filled with a PVC formulation, it is
subjected to heat and friction for 10 minutes. As a result, the gelation process will occur. The
resulting torque on the roller mixing blades is observed. Also, the temperature increase during the
gelation process is monitored. Thus, out of the Brabender experiment, a time versus torque and
temperature graph is obtained such as the one shown in Figure 24.

63
Figure 24: Typical Brabender fusion plastogram with (1) torque (Nm) versus time (min) and (2) temperature (°C) versus time
(min)

On the graph in Figure 24 some characteristic points are displayed. In point A the torque increases
very rapidly to a high value which is caused by inserting the powder into the camber. This peak is also
referred to as the loading peak. Point B represents the minimum torque of the curve at which the
powder starts to melt. At this stage of gelation, the skin around the resin grain of the PVC is torn
apart and only primary particles are present. As the temperature further increases, due to the
applied friction and shear of the rotating mixing blades, the primary particles are disappearing which
results in an increasing torque (infliction point G). The torque reaches a maximum at point X as the
interaction between the primary particles increases. The time needed to achieve the maximum value
in torque is indicated as tX and is also known as the fusion time. Subsequently, the maximum in
torque decreases as result of a strongly increasing temperature, making the ‘melt’ less viscous and
thus less torque is exerted. At point E the gelation process of the PVC is complete and a
homogenous melt is obtained. This point is often called the final torque [36][38].
Instead of a torque versus time curve, a torque versus temperature curve can be made. In Figure 25,
the torque versus temperature curve with the same characteristic points for the same measurement
as in Figure 24 is displayed. Out of these curves the same information can be obtained since the
temperature varies as function of time due to the shear and friction on the polymer. By using a
torque versus temperature curve, it is easier to compare the different results.

64
Figure 25: Typical Brabender torque (Nm) versus temperature (°C) fusion plastogram

2.2.2 Capillary rheometry


2.2.2.1 General set-up
Lubricating properties were also determined using a capillary rheometer. For these measurements
the Göttfert RG25 was used in combination with an extruder. The GöttFert LabRheo 3.7.0 software
was used in order to record the data. In Figure 26 the general set-up for these measurements is
given.

Figure 26: General set-up of the capillary rheometer in combination with the extruder

The powder was introduced into the hopper via which it is brought to the rotating screw of the
extruder. The barrel of the rheometer is simultaneously filled via an extruder. After a resting period,
the polymer melt is forced through a capillary with a diameter of 1 mm and a length of 10 mm at a
certain shear rate. A more detailed representation of the piston and capillary is given in Figure 27.

65
Figure 27: More detailed representation of the piston and capillary of the Göttfert RG25.

The temperature in the capillary and the barrel is monitored via thermocouples and controlled using
heating bands. A force sensor will measure the force on the piston in order to control height while
filling the reservoir. During the test, the polymer melt is forced through the capillary by applying a
certain force out of which the apparent shear rate is determined. This apparent or Newtonian shear
rate at the wall is determined using following equation:
4𝑄
𝛾𝑎𝑤
̇ =
𝜋 𝑅3
where 𝛾̇𝑎𝑤 represents the Newtonian shear rate at the wall, Q is the volumetric flow rate and R the
radius of the capillary. However, to obtain the real shear rate on the wall, the Weissenberg-
Rabinowitsch correction is applied since polymers are subjected to shear thinning behavior. The
following equation was used to determine the real shear rate:
1 𝑑 (ln 𝑄)
𝛾̇𝑤 = 𝛾̇𝑎𝑤 [3 + ]
4 𝑑 (ln 𝜏𝑎𝑝 )
in which 𝜏𝑎𝑝 is the apparent shear stress. The apparent shear stress is determined out of the
required pressure. This required pressure is measured using a pressure sensor. The apparent shear
stress can then be determined using the following equation:
𝑝𝑑
𝜏𝑎𝑝 = × 105 𝑃𝑎
4𝐿
with p represents the pressure of the melt in bar, d the diameter of the capillary in cm, and L the
length of the capillary in cm. Out of the apparent shear stress, the apparent viscosity (ηap) is
determined by dividing the apparent shear stress through the real shear rate (𝛾𝑤̇ ):
𝜏𝑎𝑝
𝜂𝑎𝑝 =
𝛾̇𝑤
2.2.2.2 Procedure
During all measurements, the temperature of the capillary was kept at 185 °C and the rotational
speed of the extruder screw was set at 20 rpm.
The recording of a shear rate versus shear stress or shear viscosity curve was based on three runs.
Each run consisted of two or three measurements at a different shear rate. At the beginning of each
run, the reservoir of the capillary was refilled. A waiting time was then introduced in order to ensure
the relaxation of the chains. Between each measurements of the same run a shorter stabilization

66
time was introduced for the pressure to stabilize. In Table 10, an overview of the waiting time before
each measurement and shear rates are given.
-1
Table 10: An overview of the used shear rates (s ) for each run

Run Apparent shear rate (s-1)

0.50

1 1.01

5.00

10.01

2 25.00

50.00

100.01
3
250.00

67
68
PART 3: RESULTS AND DISCUSSION

1 Development of the polymerization recipe


The influence of the temperature, initiator, emulsifier and chain transfer agent were examined. A
target value of X in specific viscosity was set which corresponds with a molecular weight of
approximately Y g/mol. In Table 37 in the attachments an overview of the adaptations in composition
that were made are given in comparison to AcrLub6, which is used as a reference for this research. A
thoroughly discussion of the influences of the different parameters of the polymerization and final
latex properties is given in the following sections. For confidentiality reasons, some concentrations
and results may be encoded with letters. All amounts are then expressed relative to the final
amount.

1.1 Preliminary results


Some research was already performed on the lubricating acrylic polymer. AcrLub6 was the final
recipe/result of the preliminary research. This was used as a base/starting point for this research. In
AcrLub6, a monomer emulsion with a drop size of 1 µm is created by homomixing M, L, purified
water and the emulsifier for five minutes at 10 000 rpm. Herein, EM17 was used as emulsifier. The
monomers were mixed in an A/B ratio. The polymerization was performed at a temperature of 70 °C.
As initiator an aqueous redox system of P4, P7, F and E/H6 was applied. Due to the high temperature,
these initiators will also decompose thermally. Table 11 gives an overview of the charges and in
which order they were added to the reactor.
With this recipe a specific viscosity of 10.6X was reached. In order to reach the target value
adaptations were made to this recipe.

Table 11: Different components and their amounts (PHR) in recipe AcrLub6

Charge Component Amount(PHR)


PW 240.00
Charge 1 (= initial charge) H16 0.05
EM17 0.30
E/H6 0.01
Charge 2 (= initiation charge) F 0.10
P7 R
M A
L B
Charge 3 PW 100.00
EM17 2.00
P4 0.75S
Charge 4 P7 0.14R
F 0.1
Charge 5
H16 0.075
Charge 6 H16 0.025
Charge 7 F 0.03
Charge 8 F 0.03

69
1.2 Influence of reactor temperature
The temperature of initial recipe (AcrLub6) is increased from 70 to 90 °C in test AcrLub10B_1. A
decrease in molecular weight is expected when increasing the temperature. As mentioned in section
1.1.3 in the literature review, this is a consequence of an increase in termination via
disproportionation. Due to a higher energy input, the high activation energies which are necessary
for termination by disproportionation can be overcome. On the other hand, an increase in
temperature promotes the decomposition of the initiator. As a result, a raise in radical concentration
should be noticed. Thus increasing the temperature should increase the formation of more and
shorter chains, which results in a lower molecular weight.
Specific viscosity measurements were performed on the polymers from the polymerization at 70 and
90 °C. The results of these measurements are shown in Table 12.

Table 12: Specific viscosity (-) results of the polymerizations at different temperatures (°C) for AcrLub6 and 10B_1

Recipe Temperature (°C) ηsp (-)


AcrLub6 70 10.57X
AcrLub10B_1 90 8.09X

As expected, the molecular weight of the acrylic lubricant drops significantly (± 20%) due to the
increase in temperature. This increase is limited to 90 °C. Because the polymerization reactions are
carried out in an aqueous medium, the increase in temperature is on one hand limited by the boiling
point of the water at atmospheric conditions. Temperatures higher than 90 °C could be possible
when working under pressure but this is not achievable with the current lab reactors.
From an industrial point of view a high temperature is not desired with respect to energy
consumption and a strict control on reactor pressure.
After the polymerizations, no scales were detected for both test AcrLub6 and AcrLub10B_1
concluding that a higher temperature does not increase the formation of scales. No difference was
noticed between the particle size distribution of AcrLub6 and AcrLub10B_1 thus the temperature
does not influence the particle size distribution. Solid content and residual monomer were not
available for AcrLub6. Therefore, no conclusions can be made regarding these topics.

1.3 Influence of initiator


The initiator concentration and type are adapted in order to achieve a higher radical flux into the
monomer drops, in which the initiation, propagation and termination all take place. A higher radical
flux will cause an increase in the amount of polymer chains and thus shorter chains should be formed
resulting in a lower molecular weight [9]. For this research two different initiators are used. On one
hand, a more polar redox initiator system is used consisting of P4, E/H6 and F. However, at
temperature above 50 °C this initiator will not only decompose via the redox system but will also
dissociate thermally. This redox initiator is partially soluble in water, causing the initiation to take
place in both the aqueous and monomer phase. On the other hand, a nonpolar thermal initiator A81
is used. Because this initiator is insoluble in water, the initiation will solely take place in the monomer
drops. In addition to both initiators, the water soluble redox initiator system with P7, E/H6 and F is
added in every polymerization to maintain a constant radical flux. Various tests were performed
using these initiators separately as well as in combination.

70
1.3.1 Influence of initiator type
In test AcrLub9 the influence of the two different initiators was tested in a polymerization reaction
with only monomer L. Test AcrLub9_1 represents the test with the redox initiator P4 and test
AcrLub9_2 is the test with A81 as initiator. During polymerization, the temperature was monitored.
The course of the temperature during the polymerization reactions is given in Figure 54 in the
attachments. In this figure, the orange temperature curve represents AcrLub9_1 (Reactor 1), the blue
temperature curve represents AcrLub9_2 (Reactor 2) and the yellow curve represents the bath
temperature. When looking at the course of the temperature for both reactions, the polymerization
with the redox initiator seems overall more exothermic in comparison to the polymerization reaction
with A81. Also, when looking at the temperature profiles both reactions seem to react to the
addition of P7. This indicates that a radical flux can be introduced using a water soluble redox system
regardless the hydrophobic character of the monomers.
The results of the specific viscosity measurements of the final samples are given in Table 13.

Table 13: Specific viscosity results (-) of AcrLub9 with the type and amount of initiator (PHR) for AcrLub9_1 and 9_2

Recipe Initiator Concentration (PHR) ηsp (-)


AcrLub9_1 P4 0.75S 7.3X
AcrLub9_2 A81 0.38Q 12.6X

The specific viscosity of the reaction with the redox initiator is approximately 40 % lower in
comparison to the specific viscosity of the reaction with the A81 initiator. This concludes that that
the recipe with the redox initiator produces a lower molecular weight than the recipe with A81.
The redox initiator probably generates a higher radical flux into the monomer drops. This difference
in radical flux can be explained by the way the radicals are generated. P4 is able to generate radicals
via thermal decomposition and via the redox system, whereas A81 only generates radicals via
thermal decomposition. The difference in radical flux can also be explained by a difference in
dissociation constant. These dissociation constants cannot be shown because of confidentiality
reasons. However, a higher dissociation constant was found for A81 in benzene, which should result
in a higher radical flux. But, because these dissociation constants were only found in benzene, they
could thus be different in a different media such as the monomer L.
The influence of this higher radical flux on the conversion rate was determined. In Table 14, the
actual conversion rates, based on the conversion of samples 1 and 3, are given for the test AcrLub9_1
and 9_2. The conversion of sample 2 was not determined.

Table 14: The actual and theoretical conversion rates (PHR/h) for AcrLub9_1 and 9_2 with the corresponding initiator and
monomer dosing time (min)

Recipe Initiator Monomer dosing Theoretical Actual conversion


time (min) conversion rate rate (PHR/h)
(PHR/h)
AcrLub9_1 P4 278 21.6 22.6
AcrLub9_2 A81 281 21.4 20.3

71
When comparing this theoretical to the actual conversion, the reactions seem to proceed at the
maximum rate, indicating that for a lower radial flux the zero-kinetics is still guaranteed.
After polymerization, scales were detected by sieving the latex. Because this experiment was
conducted with only monomer L, its homopolymer was formed. This polymer has a very low glass
transition temperature (Tg), causing the particles to stick easily together at room temperature and
form big agglomerates. Because of this phase separation takes place during the latex storage.
Therefore, a lower solid content was measured as expected. The particle size distribution was
determined but no particular difference was noticed.

1.3.2 Influence of initiator concentration


Because the use of the more polar redox initiation system resulted in a lower molecular weight, the
concentrations of the redox components were adapted in order to increase the radical flux even
more. From the temperature profile of test AcrLub9 in the attachments (Figure 54) it is seen that
both reactors react exothermic on the addition of F. This exothermic reaction indicates the formation
of radicals. Therefore, the same amount of F was dosed continuously throughout the polymerization,
instead of at once, in test reaction AcrLub10A to increase the amount of radicals throughout the
reaction. When comparing the specific viscosity results of AcrLub6 and AcrLub10A_1 (Table 15), no
particular difference is noticed, concluding that the continuous F dosing does not contribute in
lowering the molecular weight. However, in the following experiments, the continuous dosing of F is
maintained in order to avoid strong drops in temperature by adding F at once. The more constant
temperature is beneficial for the reaction.

Table 15: The obtained specific viscosities (-) with different ways of F addition for AcrLub6 and 10A_1

Recipe Dosing F ηsp (-)


AcrLub6 At once 10.57X
AcrLub10A_1 Continuous 10.17X

In test reaction AcrLub10A_2 the initial concentration of F was increased from 0.1 PHR tot 0.4 PHR in
comparison to AcrLub10A_1 with the intention to increase the initial radical concentration in the
beginning of the reaction. Table 16 shows the different concentrations and the obtained specific
viscosity results.

Table 16: Results of the specific viscosity measurements (-) and the corresponding concentration (PHR) of F in the initiation
charge for AcrLub10A_1 and 10A_2

Recipe Concentration F (PHR) ηsp (-)


AcrLub10A_1 0.1 10.17X
AcrLub10A_2 0.4 7.77X

Due to the increase of F, the specific viscosity seemed to be reduced by approximately 20%.
However, during test AcrLub10A, the dosing of F failed causing uncertainty about the amount of F
that was added in each reactor. Thus, it cannot be concluded that the decrease in specific viscosity is
due to the increase of the concentration of F in the initiation charge. Because the decrease in specific
viscosity is rather small and because the amount of F dosed in the reactors, the concentration in the

72
initiation charge was kept at 0.1 PHR. However, this test was not aborted in order to test whether
the monomer dosing time could be decreased. The results for this test are given in 1.7 in the results
and discussion section.
For the test reactions AcrLub10A_1 and AcrLub10A_2 no scales were detected after sieving the latex.
Because of the continuous dosage of F failed, only specific viscosity measurements were conducted
on the final samples for these reactions. Therefore no conclusion can be made regarding the
influence of the continuous dosing of F and the increase of the initial F on the conversion rate,
particle size distribution and the solid content.

In a next test (AcrLub11), the concentration of P4 in the monomer emulsion was increased from
0.75S to S PHR in an attempt to increase radical concentration during the reaction. The results of the
specific viscosity measurements with the corresponding amount of P4 are given in Table 17.

Table 17: The obtained specific viscosities (-) with different amounts (PHR) of P4 for AcrLub11_1 and 11_2

Recipe Initiator Concentration (PHR) ηsp (-)


AcrLub11_1 P4 0.75S 2.89X
AcrLub11_2 P4 S 2.51X

A slight drop in specific viscosity was noticed due to the increase of P4. From this can be concluded
that the impact of the increase in initiator concentration on the molecular weight is lower than
expected.
The determination of residual monomer was only performed on the final samples of AcrLub11_1 and
11_2. Therefore, no conclusion can be made regarding the conversion rate of the reaction. However,
a complete conversion was observed as there was no residual monomer was detected. For the
remaining properties such as solid content, scales and particle size distribution no differences are
observed.

The increase of the concentration of P4 only gave a small decrease in specific viscosity. Therefore, an
attempt to create a maximum radical flux in the monomer drops was made by using a combination
of P4 and A81, added to the monomer emulsion in AcrLub13. As a reference for this test, AcrLub12_2
was used since the only difference is the absence/ presence of A81. The specific viscosity results with
the concentration of P4 and A81 are given Table 18.

Table 18: The obtained specific viscosities by using a combination of the redox initiator P4 and A81 (PHR) for AcrLub12_2,
13_1 and 13_2

Recipe Concentration P4 Concentration A81 ηsp (-)


(PHR) (PHR)
AcrLub12_2 S 0 2.43X
AcrLub13_1 S 0.5Q 2.20X
AcrLub13_2 S Q 2.29X

When looking at the specific viscosity results, no particular difference is noticed due to the addition
of A81. Even in an addition level of A81 which is two times higher, the radical flux does not seem to

73
increase. This is probably because the dissociation constant of A81 is too low to create a significant
increase in the amount of radicals. Therefore, it can be concluded that the radical flux in the
monomer drop is constant.
The actual conversion rate was determined for AcrLub12_2, 13_1 and 13_2, displayed in Table 19.

Table 19: The actual and theoretical conversion rates (PHR/h) for polymerization reactions AcrLub12_2, AcrLub13_1 and
AcrLub13_2 with their corresponding dosing times (min) and used initiator system

Recipe Initiator Monomer Theoretical Actual conversion


dosing time conversion rate rate (PHR/h)
(min) (PHR/h)
AcrLub12_2 S PHR P4 99 60.6 60.8
AcrLub13_1 S PHR P4 + 0.5Q PHR A81 96 62.5 62.6
AcrLub13_2 S PHR P4 + Q PHR A81 109 55.0 55.3

The results in Table 19 indicate that the maximal conversion is reached for all reactions and no
residual monomer was left after the polymerization. This shows that introducing A81 as a second
initiator does not influence the conversion (rate) of the reaction.
For the other properties such as solid content, scales and particle size distribution no differences
were observed. Despite the small contribution of A81 to the radical flux, a combination of S PHR P4
and Q PHR A81 was kept during the next polymerization reactions.

1.4 Influence of emulsifier


The first polymerization experiments (AcrLub6 to AcrLub12_1) were conducted in a polymerization
medium with EM17 as emulsifier. For these experiments, 0.3 PHR was added in the initial charge and
2.0 PHR was used in the monomer emulsion in order to create monomer drops. In lab test
AcrLub12_2 the emulsifier type was changed to EM15. Also, no EM15 was added in the initial charge
and the emulsifier concentration in the monomer emulsion was lowered to 0.6 PHR. The emulsifier
type was changed because the use of EM15 is more feasible from an industrial point of view. The
emulsifier in the initial charge was removed because this had little to no contribution to the
polymerization reaction. Because most recipes in Kaneka Belgium involve emulsion polymerization,
emulsifier is added in the initial charge in order to create micelles and to stabilize the monomer
which is introduced in the reactor. Because in this research suspension polymerization is used and
monomer emulsion instead of pure monomer is introduced, this additional stabilization is not
necessary. The decrease concentration of emulsifier was based on calculations of the surface of the
monomer drops. Just enough emulsifier was added in order to cover the surface of monomer drops
for stabilization, which corresponds with one molecule of emulsifier per square Ångström.
By lowering the emulsifier concentration, fewer molecules will surround the monomer drop, making
the radical flux into the monomer drops more easily which should result in a decrease in molecular
weight. The results of these measurements are given in Table 20.

74
Table 20: Influence of the emulsifier type and concentration (PHR) on the specific viscosity (-) and thus the molecular
weight

Recipe Emulsifier Concentration (PHR) ηsp (-)


AcrLub12_1 EM17 2.3 2.20X
AcrLub12_2 EM15 0.6 2.43X

The results in Table 20 show that the emulsifier type and concentration do not influence the specific
viscosity, concluding that the same molecular weight is reached with EM15 as with EM17 as
emulsifier. This indicates that the radical flux in the monomer drops was not increased.
The influence on the conversion rate was determined. Results of the theoretical and actual
conversion rates for AcrLub12_1 and 12_2 are given in Table 21.

Table 21: The actual and theoretical conversion rates (PHR/h) for AcrLub12_1 and AcrLub12_2 with the corresponding
initiator and monomer dosing time (min) for different emulsifier types

Recipe Emulsifier Monomer Theoretical Average


dosing time conversion rate conversion rate
(min) (PHR/h) (PHR/h)
AcrLub12_1 EM17 143 42.0 42.3
AcrLub12_2 EM15 99 60.6 60.8

As shown in Table 21 maximum conversion rates were reached for AcrLub12_1 as well as AcrLub12_2
and almost no residual monomer was detected in the latexes. However, the conversion rate of
AcrLub12_1 is much lower than the conversion rate of AcrLub12_2 due to the higher monomer
dosing time. This higher dosing time is due to pumping of foam which is formed because of the high
concentration of EM17. This foam contains almost no monomer for the same volume as the
monomer emulsion. As a result, the conversion rate is lower purely due to an increased monomer
dosing time.
By changing the emulsifier type and concentration, no difference was notticed in solid content. When
sieving the latexes, no scales were detected for AcrLub12_1 and 12_2. However, for AcrLub12_2, a
film layer was deposited on the stirrer. This was most probably because the latex was sieved later
than usual.
When measuring the particle size distribution of the remaining latex, some differences were noticed.
In Figure 28 the volume particle size distribution is given for AcrLub12_1 (red) and AcrLub12_2
(green).

75
Figure 28: Influence of the emulsifier on the volume particle size distribution (µm) for AcrLub12_1 (red) and AcrLub12_2
(green)

The distribution of AcrLub12_2 seems to be bimodal. However, because this a volume particle size
distribution, the presence of large particles can strongly influence the distribution. Therefore, the
number distribution was calculated from the volume distribution and is given in Figure 29.

Figure 29: Influence of the emulsifier on the number particle size distribution (µm) for AcrLub12_1 (red) and AcrLub12_2
(green) measured with the Microtrac S350

According to the number particle size distribution, AcrLub12_2 does not have a bimodal distribution.
A D50 value of 0.0582 µm and 0.418 µm is determined for AcrLub12_1 and AcrLub12_2 respectively.
But, because the distribution for both test reactions is located at the lower limit of the measuring
range of the Microtrac S3500, measurements with the Nanotrac Wave ll were conducted. The
particle size distributions measured with the Nanotrac Wave ll is given in Figure 30. Here, the red
curve represents the particle size distribution of AcrLub12_1 and the green curve the particle size
distribution of AcrLub12_2.

76
Figure 30: Influence of the emulsifier on the intensity particle size distribution (µm) for AcrLub12_1 (red) and AcrLub12_2
(green) measured with the Nanotrac Wave ll

According to the intensity particle size distribution obtained from the Nanotrac Wave ll, a D50 of
0.0812 µm and 0.542 µm is determined for AcrLub12_1 and AcrLub12_2 respectively. For
AcrLub12_2 a slight difference in D50 is observed compared to the measurements with the Microtrac
S3500. This is due to the difference in intensity and number particle size distribution. This finding
confirms the reliability of the Nanotrac Wave ll measurements. For AcrLu12_1 a significant lower D50
value is found with the Nanotrac Wave ll, which could indicate a bimodal distribution.
The number particle size distribution was also calculated via the software of the Nanotrac Wave ll in
order to compare these values with the number distribution of the Microtrac S3500. The number
particle size distribution for the Nanotrac Wave ll measurements is given in Figure 31.

Figure 31: Influence of the emulsifier on the number particle size distribution (µm) for AcrLub12_1 (red) and AcrLub12_2
(green) measured with the Nanotrac Wave ll

For the number particle size distribution of AcrLub12_2 a D50 value of 0.439 which matches the value
obtained from the number particle size distribution of the Microtrac S3500 measurements. For
AcrLub12_1 a D50 of 0.05717 µm is determined.
The results of these particles size distribution measurements were confirmed with the aid of SEM

77
In Figure 33 and Figure 32 are pictures which were taken with SEM for respectively AcrLub12_1 and
12_2. Both pictures are 10 000 times amplified.

Figure 33: SEM picture of AcrLub12_1 which is 10 000 Figure 32: SEM picture of AcrLub12_2 which is 10 000
times magnified times magnified

When comparing both SEM pictures, no clear spherical polymer particles are detected for
AcrLub12_1. When looking closely at the SEM picture of AcrLub12_1, it seems like there are some
small particles which are entangled in some sort of matrix. An attempt was made to determine the
particle size of some of these particles which is however doubtful because no clear particle can be
distinguished. The remarkable structure of AcrLub12_1 can be explained by the high concentration of
emulsifier that was added during polymerization. This high concentration induced the formation of
micelles in which homopolymer of M was formed due to initiation in the aqueous phase. The
remaining monomer of L and M will polymerize in the existing monomer drops forming a polymer,
which as a lower glass transition temperature then predicted. The combination of small particles of
the homopolymer of M and big particles of a L/M polymer results in a bimodal particle size
distribution for AcrLub12_1.
For AcrLub12_2, spherical particles are clearly seen on the SEM pictures. The particles size of these
particles was also determined which resulted in a diameter of approximately 0.542 µm. This value
matches the D50 which was determined with the Microtrac S3500 and the Nanotrac Wave ll.
AcrLub12_2 therefore has a monomodal particle size distribution because no excess emulsifier was
present during polymerization. For this reason, a copolymer of M and L is to be expected.

1.5 Influence of chain transfer agent


Because the obtained specific viscosity values were still far from the target value, a chain transfer
agent was added to the polymerization medium in order to significantly lower the molecular weight.
The reaction mechanism of a chain transfer agent has already been discussed in 1.1.4 in the
literature review. The chain transfer agent that was used during this research is M1. In order to
determine at which dosage M1 needed to be added to meet the target molecular weight,
polymerization reactions with various M1 amounts were conducted. Due to the toxicity and the
unpleasant smell of M1, no or little of this monomer was allowed to remain unreacted in the
polymer. The influence of the addition of M1 on the molecular weight, the conversion rate of the
polymerization and the particle size distribution were examined.

Because in AcrLub12 the emulsifier type was changed, this section is divided into two parts. A first
series of polymerizations was conducted with EM17 as emulsifier (AcrLub10B_1 to AcrLub12_1). In
these polymerization reactions, fairly small amounts of M1 were added. In Table 22 the amounts of
chain transfer agent added to the polymerization reaction are given with the specific viscosity results

78
of the final samples. In Figure 34 the specific viscosity is plotted as function of the amount of the
chain transfer agent.

Table 22: Specific viscosity results (-) of the polymerizations with different concentration (PHR) of chain transfer agent (M1)
for AcrLub10B_1, 10B_2, 11_1 and 12_1

Recipe Chain transfer agent (M1) (PHR) ηsp (-)


AcrLub10B_1 0.0 8.09X
AcrLub10B_2 0.075U 3.94X
AcrLub11_1 0.15U 2.89X
AcrLub12_1 0.23U 2.17X

9
8
Specific viscosity *X(-)

7
6
5
4
3
2
1
0
0 0,05 0,1 0,15 0,2 0,25 0,3 0,35
Concentration M1 * U (PHR)

Figure 34: Specific viscosity (-) as function of the M1 concentration (PHR) for the recipes with EM17.

When looking at Figure 34 the specific viscosity clearly decreases with an increasing amount of M1,
concluding that much lower molecular weight could be reached using chain transfer agent. By adding
0.23U PHR of M1 to the polymerization medium, the specific viscosity has dropped to 2.17X which is
a decrease of ± 27 %. However, the influence of the addition of more M1 seemed to decrease with
an increasing amount of M1.
In order to determine if these concentrations of M1 had and influence on the conversion rate,
theoretical and actual conversion rates were determined, as displayed in Table 23.

Table 23: The actual and theoretical conversion rates (PHR/h) for polymerization reactions AcrLub10B_1, AcrLub10B_2,
AcrLub11_1 and AcrLub12_1 with their corresponding dosing times (min) and added amounts of M1 (PHR)

Monomer Theoretical Actual


Amount of M1
Recipe dosing time conversion rate conversion rate
(PHR)
(min) (PHR/h) (PHR/h)

AcrLub10B_1 0.0 111.0 54.1 52.7

AcrLub10B_2 0.075U 113.5 52.9 52.7

AcrLub11_1 0.15U 110.0 54.5 Not determined

AcrLub12_1 0.23U 143.0 42 42.3

79
The actual conversion rate was determined based on the conversion of sample 1, 2 and 3. When
comparing the theoretical and the actual conversion rate, it can be concluded that these small
concentrations of M1 did not influence the conversion rate and that complete conversion was
reached. Solid content measurements were also performed on all final samples, for which among the
samples no big differences were noticed. Moreover, no scales were detected for all reactions during
sieving of the latex. When comparing to the particle size distribution of precious reactions with EM17
as emulsifier, no difference was noticed.

The target value in molecular weight was however still not reached. This value was two times larger
than the target value. Therefore, the M1 amount was significantly increased in a second series, in
which EM15 was used as emulsifier. In AcrLub14_1 2.50U PHR M1 was used to determine whether it
was possible to reach the target value with the chain transfer agent M1. As displayed in Table 24 a
specific viscosity value of 0.71X was reached. This indicates that too much M1 was added but that
the target level could be reached with the addition of M1. Therefore, polymerizations were
conducted with U and 1.75U PHR M1 in AcrLub15_1 and AcrLub15_2 respectively in order to
determine the optimal M1 concentration. Because all these reactions were conducted with EM15 as
emulsifier, no comparison can be made with the previous polymerizations with M1. In Table 24 an
overview of the tests is given, indicating the amounts of M1 and the resulting specific viscosity. In
Figure 35 the specific viscosity is plotted as function of the amount of M1 conducted with EM15.

Table 24: Specific viscosity results (-) of the polymerizations with different concentration (PHR) of chain transfer agent (M1)
conducted with EM15 as emulsifier

Recipe Chain transfer agent (M1) (PHR) ηsp (-)


AcrLub12_2 0.23U 2.43X
AcrLub15_1 U X
AcrLub15_2 1.75U 0.83X
AcrLub14_1 2.50U 0.71X

2,5
Specific viscosity *X(-)

1,5

0,5

0
0 0,5 1 1,5 2 2,5 3
Concentration M1*U (PHR)

Figure 35: The specific viscosity (-) as function of the concentration (PHR) of M1 for the recipes with EM15

80
From Figure 35, it can be observed that (almost) no difference in molecular weight is established
when adding 1.75U or 2.50U PHR of M1. Therefore, it can be concluded that a value of 0.71X in
specific viscosity is most probably the lowest value achievable with the addition of M1. The target
value of X for the specific viscosity was reached in AcrLub15_1. Therefore, the lubricating properties
of this acrylic polymer were tested. The results of these tests are discussed in detail in section 2 of
the results and discussion.
The influence of these large amounts of M1 on the conversion rate was determined. The results of
the theoretical and actual conversion rate are given in Table 25, along with the monomer dosing time
and added M1 amount.

Table 25: The actual and theoretical conversion rates (PHR/h) for polymerization reactions AcrLub102_2, AcrLub15_1,
AcrLub15_2 and AcrLub14_1 with their corresponding dosing times (min) and added amount (PHR) of M1

Theoretical
Monomer dosing Actual conversion
Recipe Amount of M1 (PHR) conversion rate
time (min) rate (PHR/h)
(PHR/h)
AcrLub12_2 0.23U 96 60.6 60.8
AcrLub15_1 U 115 54.3 53.5
AcrLub15_2 1.75U 115 55.8 54.2
AcrLub14_1 2.50U 118 55.9 55.0

The actual conversion rate was determined based on the conversion of sample 1, 2 and 3, the same
way as for the other polymerizations. When comparing the theoretical and actual conversion rates, it
can be concluded that the large amount of M1 did not influence the conversion rate of the
polymerization and that the polymerization reached complete conversion.
For the remaining properties, such as solid content, scales and particle size distribution, no significant
changes were detected due to the addition of M1. Concluding that the use of high M1 concentrations
did not influence the particle size distribution and scale formation.

1.6 After treatment


After polymerizations, a latex was obtained consisting out of 70.95 % of water and 29.05 % of the
acrylic polymer. In order to separate the polymer from the water, a coagulation process was
developed. With this, a powder was formed which can be blended into a PVC formulation. During
this research, different coagulation processes were tested, all on AcrLub15_1.

1.6.1 Development of the procedure


In a first attempt to coagulate the latex, to 333.3 PHR latex 333.3 PHR purified water was added.
Subsequently 5.5V PHR of H8 was added to the mixture while stirring. Next, the mixture was heated
in order to precipitate the polymer particles. First the mixture was heated to 0.63T °C in order to test
this method. However, by filtering the mixture almost no particles were formed and the filtrate that
was obtained was not clear, indicating that the coagulation was not complete. The same method was
also tested where the mixture was heated to T °C and 1.88T °C but almost no particles were formed
and the filtrate was not clear.
In a second attempt to coagulate the latex, 3333.3 PHR iced water was stirred at 260 rpm. In the iced

81
water, 0.25V PHR H8 was added. The stirring speed was increased to 500 rpm before adding 333.3
PHR of the latex. Subsequently, 0.67Z PHR EM12 was added gradually to the mixture in order to
stabilize the formed particles. Afterwards, the mixture was heated to 1.5T °C with the aid of steam
and immediately after reaching this temperature, the mixture was quenched to 0.5T °C. The
quenched mixture was then filtrated over a Buchner filter, but no clear filtrate was obtained.
In a next attempt, the previous coagulation process was repeated but instead of 0.67Z PHR EM12, Z
PHR EM12 was used in order to ensure more stabilization for the formed particles. The mixture was
only heated to T °C on a heating plate while stirring. However, still no clear filtrate was obtained after
filtration of the mixture.
In a final attempt, the concentration of H8 was significantly increased in order to increase
destabilization of the latex. To stabilize the formed particles, STA12 and EM12 were used in various
concentrations. This procedure starts by adding 333.3 PHR of purified water to 333.3 PHR of latex.
While stirring, V PHR H8 was added in 333.3 PHR purified water to this mixture. Then the particles
were stabilized using EM12 or STA12. The amounts of EM12 and STA12 that were tested are given in
Table 26.

Table 26: Various amounts (PHR) of EM12 and STA12 used during the coagulation process

Coagulation EM12/STA12 H8 (PHR)


1 EM12: 3.33Z PHR V
2 EM12: Z PHR V
3 STA12: Z PHR V

These mixtures were heated to T °C while stirring. Subsequently these mixtures were filtered over a
Buchner filter. Clear filtrates were obtained for the three coagulation processes which indicate a
complete coagulation of the latex. The residues were dried overnight in a vacuum oven at room
temperature. When the residues were completely dry, granular lumps were obtained instead of fine
powders. Due to the low glass transition temperature of the polymer, the formed particles stick
together. Therefore, the residues were placed into the freezer for a couple of hours and afterwards
crashed with a mortar in order to obtain a fine powder. The obtained powders were used to
determine the influence of the coagulation process on the lubricating properties (see section 2.1.3 in
results and discussion).

1.6.2 Properties of the powder


In order to determine the influence of the molecular weight on the lubricating properties, latexes
having different molecular weights were coagulated. In order to coagulate these latexes, the
coagulation process with Z PHR EM12 and V PHR H8 was used. In Table 27, the experiments along
with the specific viscosity results of both the latexes and powder are given.

82
Table 27: Specific viscosity (-) results of the latex and the powder for AcrLub12_1, 12_2, 13_1, 15_1 and 15_2

Recipe ηsp latex(-) ηsp powder(-)

AcrLub12_1 2.17X 2.11X

AcrLub12_2 2.43X 2.29X

AcrLub13_1 2.20X 2.17X

AcrLub15_1 X 0.97X

AcrLub15_2 0.83X 0.74X

When comparing the specific viscosity results of the latex and the powder, no particular difference is
noticed. This concludes that the used coagulation process did not have an influence on the molecular
weight of the acrylic polymers.
The particle size distribution of each powder was also determined because of its importance on the
lubricating properties. For small particles it is easier to exert their lubricating properties because they
will disintegrate easier into individual polymer molecules. In Figure 36 the volume particle size
distributions of the coagulated powder are given.

Figure 36: Volume particle size distribution (µm) of the obtained powder of AcrLub12_1 (red), AcrLub12_2 (green),
AcrLub13_1 (yellow), AcrLub15_1 (blue) and AcrLub15_1 (gray)

When looking at these distributions, AcrLub12_1 seems to have the smallest particles and
AcrLub15_2 the biggest particles. To confirm these findings, the number particle size distributions
were calculated. The resulting distribution is given in Figure 37.

83
Figure 37: Number particle size distribution (µm) of the obtained powder of AcrLub12_1 (red), AcrLub12_2 (green),
AcrLub13_1 (yellow), AcrLub15_1 (blue) and AcrLub15_1 (gray)

When looking at the number particle size distribution, AcrLub12_1 clearly has the smallest particles
and AcrLub15_2 has the largest particles. Note that the particle size increases with an increasing
amount of M1 which could indicate that the addition of M1 increases the stickiness of the particles.
However, to confirm this theory, more tests are needed.
Overall it can be concluded that there is no influence of the coagulation process on the specific
viscosity. Although powder was obtained, the coagulation process needs some optimization. A large
amount of H8 is used in order to destabilize the latex, it is recommended to search for the optimum
concentrations of H8 and EM12 in order to reduce costs during producing.

1.7 Other
1.7.1 Monomer dosing rate
In AcrLub6, the monomers were dosed over 300 minutes in order to control the possible strong
exothermic reaction. In order to determine whether the monomer dosing could be increased the
heat load of the reaction was calculated. A maximum increase of 7.3 °C during the whole
polymerization was determined, resulting in a heat release of 0.024 °C/min. This released heat can
be easily removed by the water baths in the lab and cooling jackets in the production. This indicates
that the monomer dosing rate can probably be increased.
In AcrLub10A the monomer dosing rate was increased. Normally, this experiment was intended to
examine the influence of a continuous F dosage. Since the failure of this dosing, as indicated earlier,
it was decided to increase the monomer dosing rate. Out of this test reaction the influence of an
increase dosing rate on the specific viscosity was determined. Results of these measurements are
given in Table 28.

Table 28: Influence of the monomer dosing time (min) on the specific viscosity (-)

Recipe Monomer dosing time (min) ηsp (-)


AcrLub6 300 10.57X
AcrLub10A_1 150 10.17X

84
No big difference is noticed in the specific viscosity results, concluding that the increase in monomer
dosing rate did not influence the molecular weight of the acrylic polymer. There were also no
problems regarding the heat removal during the polymerization. The monomer dosing time could
thus be decreased.

1.7.2 Maron stability test


When latexes are transported in an industrial process, they undergo shear and friction due to the
pumping of the latex. In order to test the mechanical stability of the latex during transportation in an
industrial process, a Maron stability test was conducted on the latex of AcrLub15_1. The Maron test
is conducted by subjecting 100 grams of latex to shear and friction for 20 minutes by using a rotor at
1100 rpm and 55 °C. Subsequently, the latex was filtered and the residue was dried in a vacuum oven
at room temperature. The weight of the dried residue was then determined. The Maron stability was
calculated using the following equation:
𝑚𝑎𝑠𝑠 𝑑𝑟𝑖𝑒𝑑 𝑟𝑒𝑠𝑖𝑑𝑢𝑒
𝑀𝑎𝑟𝑜𝑛 𝑠𝑡𝑎𝑏𝑖𝑙𝑖𝑡𝑦 = × 100 %
𝑆𝑜𝑙𝑖𝑑 𝑐𝑜𝑛𝑡𝑒𝑛𝑡
100 × 100
For AcrLub15_1 a Maron stability of 11.79 % was determined, which is a high value. This high value
can be explained by temperature at which the measurement was performed. Because the acrylic
polymer has a glass transition temperature that is lower than 55 °C, this makes the product stickier
and form deposits more easily. Therefore, it is recommended to optimize the monomer composition
in order to obtain a higher glass transition temperature or increase the emulsifier concentration.

1.7.3 Molecular weight determination


GPC measurements were conducted on AcrLub10B_1, 11_1, 12_1, 12_2, 13_1 and 13_2. The results
of these measurements are given in Table 29 with their corresponding specific viscosity.

Table 29: The resulting molecular weight (g/mol) according to GPC measurements and the corresponding specific viscosity
(-)

Recipe ηsp (-) GPC: Molecular weight (g/mol)


AcrLub10B_1 8.09X 52.15Y
AcrLub11_1 2.89X 9.81Y
AcrLub12_1 2.17X 10.01Y
AcrLub12_2 2.43X 6.59Y
AcrLub13_1 2.20X 5.71Y
AcrLub13_2 2.29X 5.24Y
AcrLub15_1 X /

Due to problems with the apparatus, no GPC measurements were conducted on the AcrLub15_1.
Based on the result of the measurements in Table 29, the molecular weight of AcrLub15_1 can be
estimated between 0.50Y and 2.50Y g/mol. However, this estimation is only valid when there is a
linear relationship between the specific viscosity and the molecular weight. Therefore, GPC
measurements on AcrLub15_1 are recommended.

85
1.7.4 Determination of the particle size
The particle size distribution of AcrLub15_1 was determined using Mictorac S3500 and Nanotrac. Out
of these measurements, a D50 value of approximately 0.401 µm was determined according to the
number distributions. This value corresponds with the D50 value which was found in AcrLub12_2 in
section 1.4 in the results and discussions. This was expected since no changes were made in the
recipe regarding the particle size distribution.
In addition, SEM was performed on AcrLub15_1. Figure 38 and Figure 39 display SEM pictures taken
in a 1000 and 10 000 magnification respectively. A similar image as obtained with AcrLub12_2 was
expected. However, when comparing these SEM pictures to those of AcrLub12_2 in Figure 32, a clear
difference is noticed. This difference can be explained the high concentration of EM12 which forms a
‘blanket’ on the particles. The clear view in the SEM picture of AcrLub12_2 needs to be investigated
further.

Figure 38: SEM picture of AcrLub15_1 which is 1 000 Figure 39: SEM picture of AcrLub15_1 which is 1 000
times magnified times magnified

1.8 Conclusion
A recipe which produces an acrylic polymer that meets the target level of X was successfully
developed in lab test AcrLub15_1. This recipe polymerizes monomers M and L in an A/B ratio at a
temperature of 90 °C. This recipe uses a combination of the non-polar initiator A81 and a more polar
redox system with P4 and P7 as initiators. Because of the high temperature, the initiators in the
redox system will also dissociate thermally. As an emulsifier, EM15 was used because of the
monodisperse particle size distribution and the industrial applicability within Kaneka Belgium. In
order to reach the target value in molecular weight, the chain transfer agent M1 was added to the
polymerization reaction in a U PHR dosage amount.
This recipe produces a latex of the acrylic polymer with a specific viscosity of X and a solid content of
29.05 % at the maximum conversion rate of 53.5 PHR/h. The remaining latex had a D50 of
approximately 0.450 µm which was confirmed by Microtrac S3500, Nanotrac Wave ll and SEM
measurements.
The resulting latex of lab test AcrLub15_1 was successfully coagulated using V PHR H8 and Z PHR
EM12. However, optimization of the coagulation process is recommended. The coagulation process
could probably be completed with smaller amounts of H8.

86
2 Evaluation of the acrylic lubricants
Lubricating properties of the synthesized acrylic lubricants were tested using Brabender gelation and
shear viscosity measurements. A number of synthesized acrylic lubricants were tested with different
molecular weights to check the influence on the lubricating properties.

2.1 Brabender gelation


Lubricating properties of the synthesized acrylic lubricants can be determined with the use of
Brabender gelation tests, of which the set-up is described in section 2.2.1 in the materials and
methods. When adding an external lubricant to a PVC formulation, friction between the machinery
and the PVC is reduced, delaying the gelation of PVC [18][49].
A standard PVC formulation was used as a reference during these experiments: NWG ll. The
composition of this formulation is given in Table 30. In this formulation, two lubricants are present.
The first lubricant is Loxiol G-11, which is an internal lubricant consisting of partial esters of fatty
acids with glycerol. The second is Licowax E which is added for external lubrication. This lubricant
consists of esters of montanic acids. Both internal and external lubrication are needed in PVC
formulation because otherwise the processing would be very difficult due to the high melt viscosity
as already mentioned in section 3.3 in the literature review.

Table 30: Composition of the standard PVC formulation: NWG ll

Compound Amount (PHR) Function


PVC Solvin 257RF 100 PVC
Loxiol G-11 1.0 Internal lubricant
Licowax E 0.5 External lubricant
Reatinor 804S 1.5 Stabilizer
ESBO (STA-11) 1.5 Stabilizer

For the purpose of this research, four PVC formulations were prepared, in which the acrylic polymers
were added in various concentrations:
1) NWG II;
2) NWGII excluding Licowax E;
3) NWGII excluding Loxiol G-11;
4) NWGII excluding Licowax E and Loxiol G-11.
The gelation behavior of all these formulations was tested in order to determine the internal or
external lubricating properties of the acrylic polymer. Through the addition of the acrylic polymer,
the gelation was delayed in comparison to the formulation without the acrylic polymer. Evaluation of
the acrylic polymers was done based on the start of gelation, the speed of gelation and final torque.
The main goal was to create the same lubricating properties with the acrylic lubricating polymer as
with the current lubricants in the standard NWG ll formulation. In order to achieve this, the influence
of the dosing amount, the molecular weight and the coagulation process of the acrylic polymer were
evaluated.

87
2.1.1 Influence of dosing amount
For AcrLub15_1 and AcrLub15_2 the target molecular weight was reached with acceptable amounts
of M1. Therefore, the influence of the dosing amount of these components was determined because
these components were expected to have the best lubricating properties. To check whether these
components do have lubricating properties, they were first added to the standard NWG formulation
in various concentrations: 0.5, 1.0 and 5.0 PHR. Subsequently, 54 grams of these mixtures were
added in to the mixing chamber. A decrease in maximum torque and a delay in gelation were
expected compared to the standard NWG formulation. In Figure 40 and Figure 41, the gelation
curves are given for different dosing amounts of AcrLub15_1 and AcrLub15_2 respectively.
100

90

80

70
Torque (Nm)

60
Blanco-NWG
50
15_1 Z PHR EM12 - 0.5phr
40
15_1 Z PHR EM12 - 1.0 phr
30
15_1 Z PHR EM12 - 5.0phr
20

10

0
110 120 130 140 150 160 170 180
Temperature (°C)

Figure 40: Temperature (°C) versus torque (Nm) for test AcrLub15_1 (coagulated with Z PHR EM12) added in a standard
NWG formulation with 0.5, 1.0 and 5.0 phr. A chamber filling of 54 grams was used

100

90

80

70
Torque (Nm)

60

50 Blanco-NWG

40 15_2 Z PHR EM12 - 0.5phr

30 15_2 Z PHR EM12 - 5.0phr

20

10

0
110 120 130 140 150 160 170 180
Temperature (°C)

Figure 41: Temperature (°C) versus torque (Nm) for test AcrLub15_2 (coagulated with Z PHR EM12) added in a standard
NWG formulation with 0.5 and 5.0 phr. A chamber filling of 54 grams was used

88
When looking at Figure 40 and Figure 41, pretty similar gelation curves were obtained for
AcrLub15_1 and AcrLub15_2, which is probably because there is only a small difference in molecular
weight.
The maximum torque values of the gelation curves seem to decrease with an increasing amount of
acrylic polymer in the NWG formulation. A decrease in maximum torque indicates an easier rotation
of the mixing blades. This is only possible when the internal friction or/and the friction with the
machinery is reduced. Concluding that an increase in the acrylic lubricant polymer enhances internal
or/and external lubrication.
An important factor to look at is the start point of gelation. The gelation seems to occur later when
more acrylic polymer is added, indicating the reduction of friction and shear. However, when adding
5 PHR of the acrylic polymer to the formulation, the lubrication suddenly starts sooner for
AcrLub15_1 and AcrLub15_2, so far for an unidentified reason.
The rate of gelation is determined by the slope of the increasing part gelation peak. Out of the
gelation curves above it can be concluded that the gelation rate decreases as the amount of acrylic
polymer increases. This is a consequence of the increasing lubricating effects with an increasing
amount of acrylic polymer.
When looking at the end of the gelation curves, some conclusion can be made regarding the
molecular weight of the total NWG formulation. The final torque of the gelation curves seems to
decrease with an increasing amount of acrylic polymer. This is caused by a decrease of the melt
viscosity due to the increase of small polymer chains which can go to the outside of the PVC matrix
easier. Therefore, the final torque is a representation of the molecular weight of the formulation.

In order to determine the lubricating properties of the acrylic polymer, tests were conducted with
AcrLub15_1 in the NWG formulation without Loxiol G-11 and the Licowax E (formulation 4). Here, the
aim was to achieve similar gelation properties compared to the standard NWG formulation
(formulation 1). From these formulations, also 54 grams was added into the mixing chamber. Results
of these experiments are given in Figure 42.
100

90

80

70
Torque (Nm)

60 Blanco-NWG
50 Blanco-NWG w/o E and G11

40 15_1 Z PHR EM12 - 0.5phr

30 15_1 Z PHREM12 - 1.0 phr


15_1 Z PHR EM12 - 5.0phr
20

10

0
110 120 130 140 150 160 170 180
Temperature (°C)

Figure 42: Torque (Nm) versus temperature (°C) for test AcrLub15_1 (coagulated with Z PHR EM12), of which 0.5, 1.0 and
5.0 PHR was added in the NWG formulation without Loxiol G-11 and Licowax E (formulation 4), with a chamber filling of 54
grams

89
Figure 42 confirms the addition of the acrylic polymer in formulation 4 had the same influence on the
maximum torque, molecular weight, the start of gelation and the gelation rate. When looking at the
obtained gelation curves in comparison to the curve of the standard NWG formulation (formulation
1), the curve with 0.5 PHR acrylic polymer seems to start at a similar time as the standard NWG
formulation. This is caused by a similar external lubrication to Licowax E of the acrylic polymer. The
maximum torque for this formulation is however, higher than the maximum torque of the standard
NWG formulation. This is due to the lack of internal lubrication causing more internal friction which
results in a higher torque. The maximum torque of the gelation curve obtained from the formulation
with 1.0 PHR acrylic polymer seems to be a good match. The gelation starts however later compared
to the NWG formulation due to excessive external lubrication. The resulting residue of the test with
0.5 PHR acrylic lubricant was slightly sticky. The stickiness reduced with an increasing amount of the
acrylic lubricant which indicates that the acrylic polymer has external lubricating properties.
As mentioned already in 3.3 in the literature review, PVC needs external as well as internal
lubrication in order to obtain optimum processing condition. Because the acrylic lubricant is believed
to have external lubricating properties, no internal lubrication was present in the latter Brabender
experiment.

Further, Brabender experiments were conducted with 0.5 PHR and 1.0 PHR of the acrylic polymer in
a NWG formulation without the Licowax E but with Loxiol G11 as an internal lubricant (formulation
2). This experiment was conducted with a chamber filling of 54 g. The results of these Brabender
gelation tests are given in Figure 43.
100

90

80

70
Torque (Nm)

60
Blanco-NWG
50
Blanco-NWG w/o E
40
15_1 Z PHR EM12 - 0.5phr
30
15_1 Z PHR EM12 - 1.0phr
20

10

0
110 120 130 140 150 160 170 180
Temperature (°C)

Figure 43: Torque (Nm) versus temperature (°C) for test AcrLub15_1 (coagulated with Z PHR EM12), of which 0.5 and 1.0
PHR was added in the NWG formulation without Licowax E (formulation 2), with a chamber filling of 54 grams

When looking at Brabender gelation curves in Figure 43, the addition of 0.5 PHR of the acrylic
polymer to the standard NWG formulation with the internal lubricant Loxiol G11 (formulation 2)
seems to be a good match for the standard NWG formulation. By adding the internal lubricant to the
formulation, the gelation is delayed just enough in order to match the NWG formulation. The rate of
gelation, the maximal torque and the final torque are similar to the standard NWG formulation
(formulation 1). Therefore, out of these results it can be concluded that under these conditions and

90
within this PVC formulation, the external lubricant Licowax E can be replaced with the developed
acrylic polymer.

2.1.2 Comparison to other lubricants


Lubricating properties of the acrylic polymer were compared to commercially available lubricants. On
one hand, Kane Ace™ PA101 was used. This product is intended as a metal release agent for PVC
formulations and has molecular weight which is approximately ten times higher than the molecular
weight of the acrylic polymer. On the other hand, a sample of EVONIK was obtained which is an
acrylic polymer which claims to have external lubricating properties. In order to determine the
lubricating properties of these components, Brabender gelation tests were determined on a NWG
formulation without Licowax E and Loxiol G11 (formulation 4) to which 0.5 PHR of the lubricant was
added. A chamber filling of 54 grams was used for these measurements. The resulting gelation curves
are given in Figure 44.
100

90

80
Blanco-NWG
70
Blanco-NWG w/o E and
Torque (Nm)

60
G11
50 15_1 Z PHR EM12 - 0.5phr
40
Evonik - 0.5phr
30

20 Kane Ace™ PA101 - 0.5phr

10

0
110 120 130 140 150 160 170 180
Temperature (°C)

Figure 44: Torque (Nm) versus temperature(°C) for test AcrLub15_1 (coagulated with Z PHR EM12), Kane Ace™ PA101 and
the EVONIK sample of which 0.5 PHR was added in the NWG formulation without Licowax E and Loxiol G-11 (formulation 4),
with a chamber filling of 54 grams

When comparing the gelation curves in Figure 44, Kane Ace™ PA101 and the EVONIK sample seem to
start gelation sooner and have a higher maximum torque compared to the acrylic polymer. This
indicates that Kane Ace™ PA101 and the EVONIK sample are less efficient in lubricating the
formulation externally compared to the acrylic polymer. The residue obtained after the gelation test
of the EVONIK sample was very sticky which is an indication of insufficient external lubrication.
Therefore, it can be concluded that the EVONIK sample exerts mainly internal lubrication. The
residue of Kane Ace™ PA101 was less sticky in comparison to the residue of the EVONIK sample. Kane
Ace™ PA101 therefore seems to exert internal as well as some external lubrication. The residue of
the acrylic polymer was almost not sticky, which indicates good lubricating properties. When
comparing the gelation rate for the three samples, no particular difference is noticed. Thus, it can be
concluded that the acrylic polymer has the best external lubricating properties of the three samples.

91
2.1.3 Influence of molecular weight on the lubricating properties
The influence of the molecular weight on the lubricating properties was determined by testing
powders of the acrylic lubricant with various molecular weights. Powders from lab tests AcrLub12_1,
AcrLub12_2, AcrLub13_1 and AcrLub15_1 were tested. The specific viscosities for these powders,
which are an indication for the molecular weights, are given in Table 27 in section 1.6 in the results
and discussion section. The influence of the molecular weight was tested by adding 0.5 PHR of the
acrylic lubricant to the standard NWG formulation without Licowax E and Loxiol G11 (formulation 4).
A chamber filling of 54 grams was used. The results of the Brabender gelation test are given in Figure
45.

100

90

80

70
Blanco-NWG w/o E and G11
Torque (Nm)

60
Blanco-NWG
50
12_1 Z PHR EM12 - 0.5phr
40
12_2 Z PHR EM12 - 0.5phr
30
13_1 Z PHR EM12 - 0.5phr
20 15_1 Z PHR EM12 - 0.5phr
10

0
110 120 130 140 150 160 170 180
Temperature (°C)

Figure 45: Torque (Nm) versus temperature (°C) for test AcrLub12_1, 12_2, 13_1 and 15_1 (all coagulated with Z PHR EM12)
of which 0.5 PHR was added in the NWG formulation without Licowax E and Loxiol G-11 (formulation 4), with a chamber
filling of 54 grams

By examining Figure 45, gelation seems to start sooner when the molecular weight increases. This
can be explained by the shorter chains of lower molecular weight polymers which go easier to the
outside of the PVC matrix and exert external lubricating properties. Longer chains on the other hand,
increase (internal) friction, causing sooner gelation. However, differences between the start of the
gelation are small. The maximum torque is also lower when the molecular weight decreases, because
of an easier rotation of the mixing blades. This indicates that lower molecular weight acrylic polymers
exert more lubrication and thus less friction. No differences in gelation rate are observed in these
gelation curves. The gelation peaks for AcrLub12_1, AcrLub12_2 and AcrLub13_1 are very similar
because the molecular weight is almost the same for these powders and exert therefore almost the
same lubricating behavior. Also, almost no difference is noticed in final torque in comparison to the
NWG without acrylic polymer. This is because only small amounts of the acrylic polymer were added
which do not significantly change the molecular weight of the total formulation.
The same experiments were conducted with a standard NWG formulation without the Licowax E but
with the Loxiol G11 as an internal lubricant (formulation 2). The results of these Brabender gelation
tests are given in Figure 46.

92
100

90

80

70
Blanco-NWG
Torque (Nm)

60
Blanco-NWG w/o E
50
12_1 Z PHR EM12 - 0.5phr
40
12_2 Z PHR EM12 - 0.5phr
30
13_1 Z PHR EM12 - 0.5phr
20 15_1 Z PHR EM12 - 0.5phr
10

0
110 120 130 140 150 160 170 180
Temperature (°C)

Figure 46: Torque (Nm) versus temperature (°C) for test AcrLub12_1, 12_2, 13_1 and 15_1 (all coagulated with Z PHR EM12)
of which 0.5 PHR was added in the NWG formulation without Licowax E (formulation 2), with a chamber filling of 54 grams

When looking at Figure 46 the gelation of all formulations with the acrylic lubricants seems to start
almost at the same time, which indicates that due to the addition of the internal lubricant the
difference in external lubricating properties are less distinct. For the maximum torque the same
conclusion can be made as with the gelation test conducted without any lubricants, displayed in
Figure 45. The maximum torque decreases as the molecular weight decreases. The rate of gelation
however, seems also to decrease with a decreasing molecular weight, indicating better lubrication
when using lower molecular weight acrylic polymers.
However, when comparing Figure 43 and Figure 46, it can be noticed that the addition of AcrLub15_1
to formulation 4 does not match the standard NWG compound one to one for both cases. The
difference between the tests can be related to the day that the both tests were conducted. Small
variations in humidity level can be sufficient in order to cause the observed differences, because of
the hygroscopic character of PVC formulations. In addition, the differences between the gelation
curves are small and look different than expected.
Therefore, it is recommended to repeat the same Brabender experiments with a chamber filling of 50
grams in order to spread out the gelation curves more. It is also recommended to dry the samples
after mixing to obtain more reproducible results.

2.1.4 Influence of the coagulation process


The influence of the coagulation process was tested on the gelation behavior. As explained in 1.6 in
the results and discussion section, different coagulation processes were tested in order to obtain
powder. In a first coagulation process, the particles are stabilized using EM12. Coagulation processes
with Z and 3.33Z PHR EM12 were conducted on AcrLub15_1. Brabender gelation test were
performed using a standard NWG formulation without Licowax E and Loxiol G11 (formulation 4) and
various amounts of the acrylic lubricant. A chamber filling of 54 grams was used. The obtained
gelation curves of these Brabender gelation tests are given in Figure 47.

93
100

90 15_1 Z PHR EM12 - 0.5phr


80
15_1 3,33Z PHR EM12 -
70 0.5phr
Torque (Nm)

60 15_1 Z PHREM12 - 1.0 phr

50 15_1 3,33Z PHR EM12 -


40 1.0phr
15_1 Z PHR EM12 - 5.0phr
30

20 15_1 3,33Z PHR EM12 -


5.0phr
10

0
110 120 130 140 150 160 170 180
Temperature (°C)

Figure 47: Torque (Nm) versus temperature (°C) for test AcrLub15_1 (coagulated with Z and 3.33Z PHR EM12) of which 0.5
PHR was added in the NWG formulation without Licowax E and Loxiol G-11 (formulation 4), with a chamber filling of 54
grams

When looking at Figure 47, the gelation starts a little bit sooner when using Z PHR of EM12 in
comparison to 3.33Z PHR EM12 in the coagulation process. The gelation occurs later with 3.33Z PHR
because EM12 exerts some lubricating properties as well. Overall the torque seems slightly higher for
the gelation curves with Z compared to the 3.33Z PHR EM12, which is also due to the lubricating
properties of EM12. For the rate of gelation no differences are noticed between Z and 3.33Z PHR
EM12. Out of these Brabender gelation curves it can be concluded that using more EM12 during the
coagulation process enhances the lubricating properties of the acrylic polymer, but the differences
are very small. Therefore, the coagulation process with Z PHR EM12 seems more interesting.
In a second coagulation process, Z PHR STA12 was used in order to stabilize the formed particles. A
standard NWG formulation without Licowax E and Loxiol G11 (formulation 4) was used and different
amounts of the acrylic polymer were added. A chamber filling of 50 grams was used, which should
spread out the gelation curves more since the friction and shear is lowered. The results of the
Brabender gelation tests are given in Figure 48.

94
100

90
15_2 Z PHR EM12 -
80 0.5phr

70 15_2 Z PHR STA12 -


0.5phr
Torque (Nm)

60
15_2 Z PHR EM12 -
50 0.75phr

40 15_2 Z PHR STA12 -


0.75phr
30
15_2 Z PHR EM12 -
20 1.0phr

10 15_2 Z PHR STA12 -


1.0phr
0
110 120 130 140 150 160 170 180
Temperature (°C)

Figure 48: Torque (Nm) versus temperature (°C) for test AcrLub15_1 (coagulated with Z PHR EM12 and Z PHR STA12) of
which 0.5 PHR was added in the NWG formulation without Licowax E and Loxiol G-11 (formulation 4), with a chamber filling
of 50 grams

When looking at Figure 48, the gelation of the NWG formulation with the acrylic lubricant coagulated
with STA12 seems to start later in comparison to the gelation of the NWG formulation with the
acrylic lubricant which was coagulated with EM12. This shows that STA12 has better lubricating
properties than EM12. The difference in lubricating properties between EM12 and STA12 can be
explained by the structure. The structures of EM12 and STA12 cannot be shown because of
confidentiality reasons. STA12 has however a very voluminous structure with large side groups in
comparison to EM12. Due to this large structure, the incompatibility with the PVC matrix increases
which enhance the external lubricating properties. No difference is noticed in the rate of gelation or
in maximum torque.

2.1.5 Other experiments


As already mentioned in 2.1.1 in the results and discussions section, the NWG formulation with only
the acrylic polymer (formulation 4) is in need of an internal lubricant in order to obtain similar
behavior as the standard NWG formulation. Therefore, Kane AceTM PA310 was added to formulation
4 in order to improve internal lubrication. Kane AceTM PA310 is a high molecular weight acrylic
polymer which is used as a processing aid. In order to determine the lubricating properties of the
formulation, different amounts of acrylic polymer and Kane AceTM PA310 were added to a NWG
formulation without Licowax E and Loxiol G11 (formulation 4). These tests were carried out using
AcrLub15_2, since its comparable behavior compared to AcrLub15 and the limited sample of
AcrLub15_1. The Brabender gelation tests were carried out with a chamber filling of 50 g. The
resulting gelation curves are given in Figure 49.

95
100

90

80

70 Blanco-NWG
Torque (Nm)

60
Blanco - NWG w/o E and G-
11
50
0,5 phr 15_2 Z PHR EM12 +
40 1,0 phr Kane AceTM PA310
30 1,0 phr 15_2 Z PHR EM12 +
1,0 phr Kane AceTM PA310
20
1,0 phr 15_2 Z PHR EM12 +
10 2,0 phr Kane AceTM PA310

0
110 120 130 140 150 160 170 180
Temperature (°C)

TM
Figure 49: Torque (Nm) versus temperature(°C) for test AcrLub15_1 (coagulated with Z EM12) and Kane Ace PA310 was
added in the NWG formulation without Licowax E and Loxiol G-11 (formulation 4), with a chamber filling of 54 grams

When looking at the obtained gelation curves, the gelation of the formulation with 1.0 PHR of the
acrylic polymer and 1.0 PHR of the Kane AceTM PA310 seems to start at the same time as the gelation
of a standard NWG formulation. The gelation rate and the maximum torque are however higher
when comparing to the standard NWG formulation. These observations can possibly be related to
the molecular weight of the processing aid, which is apparently higher compared to the Loxiol G-11.
This way, a higher internal friction is present, by which the faster gelation and higher torque can be
explained.
The combination of Kane AceTM PA310 and the acrylic lubricant are interesting for Kaneka from an
industrial point of view regarding the after treatment of the latex. The acrylic polymer has a low glass
transition temperature, making spray drying of the latex impossible. By blending the latexes of Kane
AceTM PA310 and the acrylic polymer, the glass transition temperature increases making it possible to
spray dry the blended latex.

2.2 Shear viscosity measurements (Contifeed)


A mentioned in section 3 in the literature review, lubricants can exert their function by lowering the
melt viscosity. As a consequence, lubricating properties can be defined by shear viscosity
measurements. In order to get an idea of the shear viscosity of the NWG formulations with and
without lubricants, shear viscosity tests were conducted on the standard NWG formulation
(formulation 1) , NWG without Licowax E (formulation 2), NWG without Loxiol G11 (formulation 3)
and NWG without both Licowax E and Loxiol G11 (formulation 4). The composition of the standard
NWG formulation was already given in section 2.2.1 (Table 30) in the results and discussion section.

2.2.1 Reproducibility
As mentioned in 2.2.2 in the materials and methods section, the shear rate is applied from which the
pressure is determined in order to force the polymer melt through the capillary. To get an idea of the
reproducibility of the experiment, measurements with the standard NWG formulation (formulation
1) and with the NWG formulation without the Licowax E (formulation 2) were repeated twice. In

96
Table 31 the resulting pressures for both measurements on the NWG formulation are given along
with the difference in pressure and the accompanied percentage error.
-1
Table 31: The calculated pressures (bar) and the applied shear rate (s ) for both shear viscosity measurements on the
standard NWG formulation along with the pressure differences and percentage error

Applied Pressure
Measurement 1: Measurement 2: Percentage
shear rate difference
Pressure (bar) Pressure (bar) error (%)
(s-1) (bar)

0.50 13.96 14.79 0.83 5.95

1.01 20.49 18.80 1.69 8.27

5.00 35.02 36.20 1.19 3.39

10.01 49.24 51.71 2.47 5.02

25.00 75.24 76.49 1.25 1.66

50.00 99.58 103.77 4.19 4.21

100.01 130.60 132.43 1.83 1.40

250.00 173.91 176.67 2.76 1.59

The pressure difference between the two measurements seems to vary, but overall a pressure
deviation of 1 to 2 bar appears to be valid. For the first two measurements with a shear rate of 0.50
and 1.01 s-1, this is a fairly large deviation on the applied pressure. Therefore, the reliability of the
first two measurements is doubtful.
The same thing was done for the repeated measurement on the NWG formulation without Licowax E
(formulation 2). The pressure measurements for both experiments on the NWG formulation without
Licowax E (formulation 2) are given along with the pressure difference and the percentage error.
1
Table 32: The calculated pressures (bar) and the applied shear rate (s ) for both shear viscosity measurements on the NWG
formulation without Licowax E along with the pressure differences and percentage error

Applied Pressure
Measurement 1: Measurement 2: Percentage
shear rate difference
Pressure (bar) Pressure (bar) error (%)
(s-1) (bar)

0.50 15.78 16.73 0.96 6.06

1.01 20.02 20.82 0.79 3.97

5.00 38.49 39.19 0.70 1.82

10.01 53.57 53.01 0.56 1.05

25.00 79.74 81.10 1.36 1.71

50.00 106.10 106.36 0.26 0.24

100.01 137.91 137.14 0.76 0.55

250.00 182.77 181.53 1.25 0.68

The same conclusions can be made when looking at the results from the measurements on the NWG
formulation without Licowax E as for the measurements on the standard NWG formulation. For
these measurements overall a difference in pressure of 1 bar was noticed and seems to be more
stable compared to the pressure differences for the NWG formulation. From these measurements it

97
is clear that the percentage error decreases as the pressure and shear rate increase. At higher shear
rates, the contribution of the pressure difference is small/negligible. This decrease in percentage
error can be explained by the range of the pressure sensor is limited to 1500 bar. When measuring
lower pressures, the fault on the measurement will increase.
The calculated shear viscosities for the repeated tests were also compared in order to get an idea of
the error on the final results. In order to be able to compare the specific viscosities, a fit was made
for all measurements using the Carreau - Yasuda model via which the specific viscosity for every
shear rate can be determined. These calculations were done by using the following equation:
𝜂 − 𝜂∞
= [1 + (𝜆𝛾̇ )𝑎 ](𝑛−1)/𝑎
𝜂0 − 𝜂∞
in which η0 represents the asymtotic viscosity at zero shear rate and no yield stress, η∞ the asymtotic
viscosity at infinite shear rate and no yield stress, λ is a time constant, 𝛾̇ the applied shear rate, n is
the power-law index, a is the width of the transition region between η0 and the start of the power-
law region [50]. In Figure 50 the measuring points and the fitted data are plotted for the two
measurements on the standard NWG formulation.
1,00E+05
Shear viscosity (Pa.s)

1,00E+04

NWG
NWG fit
NWG repeat
1,00E+03
NWG repeat fit

1,00E+02
1,00E-01 1,00E+00 1,00E+01 1,00E+02 1,00E+03 1,00E+04
Shear rate (s-1)

-1
Figure 50: Shear viscosity (Pa.s) as function of shear rate (s ) for the repeated measurements on the standard NWG
formulation, including data points fitted with the Yasuda model

In Table 33 the shear rates and the calculated shear viscosity from the fitted data are given along
with the percentage error between the two measurements.

98
1
Table 33: Comparison of the shear viscosities (Pa.s) for the same shear rate (s ) for the standard NWG formulation along
with the percentage error between the measurements

(absolute)
Applied Measurement 1: Measurement 2: Shear
Percentage error
shear rate Shear viscosity Shear viscosity viscosity
(%)
(s-1) (Pa.s) (Pa.s) difference
(Pa.s)

0.5 65332.55 65526.76 194.21 0.30%

1 43738.49 43477.64 260.85 0.60%

5 16711.52 16772.88 61.36 0.37%

10 11037.22 11128.96 91.74 0.83%

50 4212.58 4293.35 80.77 1.92%

100 2782.22 2848.68 66.46 2.39%

500 1061.89 1098.97 37.08 3.49%

1000 701.33 729.18 27.85 3.97%

From the measurements with the standard NWG formulation, the percentage error seems to
increase with an increasing shear rate. From these measurements it can be concluded that an error
of ± 4.0 % is possible and must be taken into account when reviewing the results of the shear
viscosity measurements.
The same thing was done for the measurements with the NWG formulation without Licowax E
(formulation 2). In Figure 51 the measuring points and the fitted data are plotted for the two
measurements on the NWG formulation without Licowax E.
1,00E+05
Shear viscosity (Pa s)

1,00E+04

NWG w/o Licowax E

NWG w/o Licowax E

1,00E+03
NWG w/o Licowax E
repeat
NWG w/o Licowax E
repeat fit

1,00E+02
1,00E-01 1,00E+00 1,00E+01 1,00E+02 1,00E+03 1,00E+04
Shear rate (s -1)

-1
Figure 51: Shear viscosity (Pa.s) as function of shear rate (s ) for the repeated measurements on the NWG formulation
without Licowax E (formulation 2), including data points fitted with the Yasuda model

99
As for the measurements with the standard NWG formulation, the fitted data show a good
correspondence to the measured data. As a result, shear viscosities can also be calculated for the
desired shear rates using the fitted data for these measurements. In Table 34, shear rates with the
calculated shear viscosity from the fitted date are given together with the percentage error between
the two measurements.
-1
Table 34: Comparison of the shear viscosities (Pa.s) for the same shear rate(s ) for the NWG formulation without Licowax E
along with the percentage error between the measurements

(absolute)
Applied Measurement 1: Measurement 2: Shear
Percentage error
shear rate Shear viscosity Shear viscosity viscosity
(%)
(s-1) (Pa.s) (Pa.s) difference
(Pa.s)

0.5 69068.90 73047.75 3978.85 5.76%

1 45725.68 47910.88 2185.20 4.78%

5 17548.88 17993.63 444.75 2.53%

10 11617.87 11801.72 183.85 1.58%

50 4458.78 4432.31 26.47 0.59%

100 2951.84 2907.08 44.76 1.52%

500 1132.88 1091.80 41.08 3.63%

1000 750.00 716.09 33.91 4.52%

Overall the percentage error seems slightly higher in comparison to the measurement with the
standard NWG formulation. The second measurement with the NWG formulation without Licowax E
was done at the end. Therefore, it is possible that some deposits have been formed throughout all
the other measurements. In order to avoid this problem and obtain more reproducible results, the
capillary should be cleaned every other measurement. The absence of the external lubricant could be
another reason for the larger error on the results. This way, a higher friction between the melt and
the capillary exists, creating a less smooth and thus stable flow. Overall, these measurements show
that a percentage error of ± 6.0 % is possible during the measurements and should therefore be
taken into account.

2.2.2 Influence of ‘standard’ lubricants on shear viscosity


In order to determine the lubrication behavior for the four different NWG formulations (formulations
1-4), shear viscosity measurements were conducted. For all measurements a fit was made using the
Carreau - Yasuda model. In Figure 52 the shear viscosity is plotted as function of the shear rate. In
A duplicate of this figure was added into the attachments (Figure 55) for a much clearer view on the
location of the curves.

100
7,00E+04
NWG w/o E and G11
NWG w/o E and G11 fit
NWG w/o G11
NWG w/o G11 fit
NWG w/o Licowax E
Shear viscosity (Pa s)

NWG w/o Licowax E fit


NWG
NWG fit

7,00E+03

7,00E+02
1,00E+00 1,00E+01 1,00E+02 1,00E+03
Shear rate (s -1)

-1
Figure 52: Shear viscosity (Pa.s) versus shear rate (s ) for the measured and fitted date for the different NWG formulations
with adjusted axes

To compare the different NWG formulations, the percentage difference was determined relative to
the standard NWG formulation. The results are given in Table 35.

Table 35: Percentage difference (%) of the shear viscosities of all NWG formulation (formulation 2-4) relative to the
-1
standard NWG formulation (formulation 1) accompanied with the applied shear rate (s )

Shear viscosity Shear viscosity Shear viscosity percentage


Applied shear
percentage difference percentage difference difference NWG w/o G-11
rate (s-1)
NWG w/o E NWG w/o G-11 and E
0.5 5.72% 11.58% 13.95%
1 4.54% 10.18% 12.54%
5 5.01% 10.28% 12.71%
10 5.26% 10.37% 12.83%
50 5.84% 10.59% 13.12%
100 6.10% 10.69% 13.24%
500 6.69% 10.91% 13.53%
1000 6.94% 11.01% 13.65%

Since for the formulation without G11 (formulation 3) and the formulation without G11 and E
(formulation 4) the percentage difference is higher compared to the possible percentage error (± 6
%), the difference between these measurements is significant. The formulation without Licowax E
(formulation 2) on the other hand, the percentage differences (± 6 %) lay in the range of the
percentage error of ± 6 %. Therefore, conclusions regarding this formulation should be re-examined.
When looking closely at the plot and the percentage differences, the standard NWG formulation

101
(formulation 1) has overall the lowest shear viscosity out of all the NWG formulations. The NWG
formulation without Licowax E and Loxiol G11 (formulation 4) however, has overall the highest
viscosity. These findings confirm the expectations that lubricants lower the shear viscosity by
reducing internal and external friction.
When looking at the plots of the NWG formulation without Licowax E (formulation 2) and the NWG
formulation without Loxiol G11 (formulation 3), the formulation without the Licowax E seems to
reduce the shear viscosity more. Because Loxiol G11 is an internal lubricant, it will be present in the
full PVC matrix. On the other hand Licowax E, which is an external lubricant, will go to the outside of
the PVC matrix and exerts there its lubricating properties. As a consequence, the Loxiol G11 can exert
its lubricating properties over a much greater area in comparison to the Licowax E, which causes the
difference in shear viscosity between the two formulations.
However, the findings regarding the NWG formulation without the Licowax E (formulation 2) are
under reservation because of the insignificant percentage difference with the standard NWG
formulation. Therefore, it is recommended to repeat these tests with different processing conditions.
For example, a capillary with a smaller diameter can be used or the temperature of the capillary can
be decreased.

2.2.3 Influence of the acrylic polymer on shear viscosity


Out of the Brabender gelation tests seemed that the addition of 0.5 PHR of the acrylic lubricant in the
NWG formulation without Licowax E (formulation 2) resulted in similar lubricating properties
compared to the standard NWG formulation. In order to confirm these results, shear viscosity
measurements were conducted on this formulation. In addition, shear viscosity measurements were
conducted with 0.5 PHR of the acrylic polymer and 1.0 PHR Kane AceTM PA310 in a NWG formulation
without the standard lubricants (formulation 4). For all measurements a fit was made using the
Carreau - Yasuda model. The measured and fitted shear viscosities as function of the shear rate are
given in Figure 53.

1,00E+05

1,00E+04
Shear viscosity (Pa s)

1,00E+03

1,00E+02

1,00E+01

1,00E+00
1,00E-02 1,00E-01 1,00E+00 1,00E+01 1,00E+02 1,00E+03 1,00E+04
Shear rate (s -1)
NWG NWG fit
NWG w/o E with G11 + AcrLub15_1 NWG w/o E with G11 + AcrLub15_1 fit
NWG w/o E + G11 and with AcrLub15_1+PA310 NWG w/o E + G11 and with AcrLub15_1 + PA310 fit

-1
Figure 53: Shear viscosity (Pa.s) versus shear rate (s ) for the measured and fitted data for a NWG formulation without
Licowax E (formulation 2) and 0.5 PHR AcrLub15_1 and for a NWG formulation without the standard lubricants (formulation
TM
4) with 0.5 PHR AcrLub15_1 and 1.0 PHR Kane Ace PA310

102
To compare the formulations to which the acrylic polymers were added, the percentage difference
was determined relative to the standard NWG formulation (formulation 1). The results are given in
Table 36.

Table 36: Percentage difference (%) of the shear viscosities of all NWG formulation (formulation 2-4) relative to the
-1
standard NWG formulation (formulation 1) accompanied with the applied shear rate (s )

Shear viscosity percentage


Applied shear Shear viscosity percentage difference NWG
difference NWG w/o E +
rate (s-1) w/o E and G-11 with AcrLub15_1 and PA310
AcrLub15_1
0.5 7.91% 3.01%
1 6.09% 4.23%
5 1.17% 10.42%
10 0.09% 13.26%
50 0.96% 20.12%
100 0.52% 23.20%
500 2.16% 30.66%
1000 3.92% 34.02%

According to the plot and the percentage differences from the NWG formulation without Licowax E
and with AcrLub15_1 seems to have a similar shear viscosity (± 3 % difference) to the standard NWG
formulation. However, there is insignificant difference (± 6 %) between the standard NWG
formulation (formulation 1) and the NWG formulation excluding Licowax E (formulation 2)
compared to the percentage error (± 6 %) on the measurement as mentioned earlier. Therefore, it is
unclear whether these findings are due to the addition of the acrylic lubricant or due to errors on the
measurements. For now, it can only be concluded that the formulation with the acrylic polymer
follows the same trend as the standard NWG formulation. Nevertheless, the same was already
concluded during Brabender gelation tests. According to this test, the acrylic polymer could be a
replacement for Licowax E in a NWG formulation. It is however recommended to repeat this test at
different processing conditions.
The plot for the NWG formulation without the standard lubricant to which the acrylic polymer and
Kane AceTM PA310 was added has a significant higher (± 17 %) shear viscosity compared to the
standard NWG formulation and the NWG formulation without Licowax E to which the acrylic polymer
and Loxiol G11 were added. These findings confirm the findings from the Brabender gelation test in
2.1.5 in the results and discussion section. This can be explained by the high molecular weight of the
Kane AceTM PA310.
As mentioned in the previous section, it is recommended to repeat these tests with other processing
conditions in order to spread the plots out more.

2.3 Conclusion
The acrylic polymer has proven to exert mainly external lubricating properties from Brabender
gelation and shear viscosity measurements. During Brabender gelation measurements, the addition
of 0.5 PHR acrylic polymers to an NWG formulation without standard lubricants displayed a similar
start in gelation but a higher maximum torque. This indicates similar external lubrication behavior for

103
Licowax E and the acrylic polymer but that there is need for internal lubrication. Brabender gelation
measurements where 0.5 PHR of the acrylic polymer was added in a NWG formulation without
Licowax E showed almost identical lubricating properties as the standard NWG formulation. The
similarity of this formulation with the standard NWG formulation was confirmed by shear viscosity
measurements. Out of these shear viscosity measurements a similar trend of shear viscosity as
function of shear rate was observed. From these findings, it can be concluded that the acrylic
polymer exerts similar lubrication as the Licowax E, at the tested processing conditions.
When comparing the acrylic polymer to the lubricant of EVONIK and Kane AceTM PA101, the acrylic
polymer seemed to have a later gelation and a lower maximal torque. Therefore, it can be concluded
that the acrylic polymer expresses better external lubricating properties.
The gelation behavior and thus the lubricating properties appeared to be influenced by the molecular
weight of the acrylic polymer. A decrease in molecular weight delayed the gelation and lowered the
maximum torque, which hints better lubricating properties for low molecular weight acrylic
polymers. However, these findings need confirmation. It is therefore recommended to repeat these
tests with a chamber filling of 50 gram to spread out the gelation curves. Additionally, the
formulations can be dried after mixing to increase reproducibility.
A clear influence of the coagulation process on the lubricating properties was also observed during
Brabender gelation tests. It showed that an increasing amount of EM12 delayed the gelation and
decreased the maximum torque. The delay in gelation and the decrease in maximum torque were
even more present by using STA12. This concludes that the use of EM12 as well as the use of STA12
enhances the lubricating properties.
According to Brabender gelation measurements, a similar start in gelation was observed compared
to the standard NWG formulation was achieved by adding 1.0 PHR of the acrylic polymer and 1.0 PHR
of Kane AceTM PA310 to a NWG formulation without standard lubricants. Compared to the standard
NWG formulation, a higher gelation rate and maximum torque were however noticed. This can be
related to the high molecular weight of Kane AceTM PA310 causing more internal friction. These
findings were confirmed by the shear viscosity measurements since it was proved that the addition
of Kane AceTM PA310 increased (± 17%) the shear viscosity compared to a standard NWG
formulation.
Overall it can be concluded that an acrylic polymer with similar external lubricating properties as the
Licowax E is produced by which it could be replaced eventually.

104
CONCLUSION
The main goal of this master’s thesis was to develop a recipe that produces an acrylic external
lubricant that possibly solves the plate-out issue from which current lubricants are suffering. The
most important characteristic of the acrylic polymer is the low molecular weight. Therefore, various
parameters of a free radical suspension polymerization process were adapted until a low molecular
weight acrylic polymer was achieved.
The recipe polymerizes the monomers M and L in an A/B ratio. A first parameter that was adapted
was the reactor temperature. The optimal temperature for this recipe was set at 90 °C.
Subsequently, the initiator concentration and type were adapted. A combination of Q PHR of the
non-polar initiator A81 and the more polar redox system S PHR P4 and R PHR P7 seemed the best
choice. As a next step, the emulsifier type and concentration was reviewed. 0.6 PHR of EM15 seemed
the best fit. In order to reach the target molecular weight, the addition of a chain transfer agent was
needed. The polymerization tests proved that with an addition of U PHR M1 the target molecular
weight was reached.
The resulting recipe produces an acrylic polymer which has a specific viscosity of X at a maximum
conversion rate of 53.5 PHR/h. The remaining latex has a solid content of 29.05 % and the polymer
particles in the latex have a D50 of approximately 0.450 µm.
In order to evaluate the lubricating properties of the acrylic polymer, a powder was obtained from
the latex. Therefore, a coagulation process was developed. This coagulation process uses V PHR H8
and Z PHR EM12 at a temperature of T °C. The coagulation process however needs some
optimization regarding the added amount of H8 and the particle size of the remaining powder.

The lubricating properties of the acrylic polymer were determined using Brabender gelation and
shear viscosity measurements. The addition of 0.5 PHR acrylic polymer to an NWG formulation with
Loxiol G-11 displayed similar gelation behavior compared to the standard NWG formulation. These
findings were confirmed by shear viscosity measurements. Here, the trend of the shear viscosity as
function of the shear rate was similar to the trend of the standard NWG formulation. This indicates
that the acrylic polymer exerts similar lubricating properties as Licowax E, under the tested
conditions, an is in need of internal lubrication.
Therefore, Kane AceTM PA310 was added to a NWG formulation without standard lubricants along
with the acrylic polymer to improve internal lubrication. A similar start of gelation was observed by
adding 1.0 PHR acrylic polymer and 1.0 PHR Kane AceTM PA310 to the NWG formulation without
standard lubricants. However, the rate of gelation and maximum torque were higher compared to
the standard NWG formulation due the high molecular weight of Kane AceTM PA310. The shear
viscosity measurements confirmed these findings. Here, the shear viscosity in function of the shear
rate was approximately 17 % higher compared to the standard NWG formulation due to the addition
of 1.0 PHR Kane AceTM PA310.
The molecular weight of the acrylic polymer appeared to influence the gelation behavior and thus
the lubricating properties. Decreasing the molecular weight delayed the gelation and lowered the
maximum torque. This indicates better lubricating properties for acrylic polymers with low molecular
weights. A repetition of these tests with a chamber filling of 50 gram and after drying of the
formulation mixture is however recommended.
The emulsifier used to stabilize the particles (EM12 and STA12) during the coagulation seemed to
enhance the lubricating properties. Overall a delay in the start of gelation and a decrease in

105
maximum torque were observed when increasing the amount. This influence appeared to be more
pronounced when using STA12 compared to EM12.

However, some additional qualitative evaluation of the acrylic polymer is necessary in the further
development of the acrylic lubricant. An evaluation of the acrylic polymer in different matrices which
are more relevant to the industrial field of application should be performed. This way, a clearer view
can be obtained of which parameters of the polymerization process should be further adapted. In
addition, an evaluation test which determines the formation of (die) plate-out should be considered.
This could be achieved by carrying out a Haake extrusion test, in which a qualitative analysis is made
of components migrating from the PVC matrix and depositing on plates. This makes it possible to
compare the plate-out formation of the acrylic polymer to other lubricants.
Subsequently, based on the evaluation results, the polymerization recipe can be adapted to improve
final properties of the acrylic polymer. Parameters such as the monomer types and ratio can be
changed. Furthermore, the structure (branched or linear) and the molecular weight distribution of
the polymer can be adapted depending on depending on the desired final properties.
Some adaptations to the polymerization recipe need also to be made regarding the industrial
application of the acrylic polymer. A first parameter that should be changed is the solid content of
the latex. The current recipe produces a latex which has a solid content of 29.05 % which is not
interesting from an industrial and economical point of view. Therefore, the solid content should be
increased to an optimal value of approximately 40%. A second parameter which can be optimized is
the concentration of M1 in the polymerization recipe. In the current recipe a relatively large amount
of M1 is used. This is not desired because of the toxicity and bad smell of M1.
Lastly, the after treatment process should be revised. Because of the low glass transition
temperature of the polymer, spray drying of the latex is not possible. Therefore, a further
optimization of the coagulation process is needed. Another possibility is to blend the latex of the
acrylic lubricant with the latex of a polymer which has a high glass transition temperature and
enhances the lubricating properties.

106
Bibliography
[1] W. V. Titow, PVC Technology. Londen, New York: Elsevier Applied Science, 1984.
[2] G. Odian, Principles of polymerization. New Jersey: John Wiley & Sons, 2004.
[3] P. Eyerer, M. Weller, and C. Hübner, Polymers – Opportunities and Risks II. Berlin Heidelberg:
Springer, 2008.
[4] R. Spiekermann, “New lubricants offer higher efficiency in PVC extrusion,” Plast. Addit.
Compd., vol. 10, no. 5, pp. 26–31, 2008.
[5] J. I. Wrozina, P. E. Burdick, and P. J. Albee, “Low molecular weight copolymer salts as lubricant
in plastic,” 4,412,040, 1981.
[6] M. Gilbert, R. Haberleitner, M. Schiller, K. Van Soom, and N. Varshney, “Plate-out – more than
just a phenomenon? Part 1: Plate- out in the extruder.,” Int. Polym. Sci. Technol., vol. 31, no.
12, pp. 3–7, 2004.
[7] N. Varshney, M. Gilbert, M. Walon, and M. Schiller, “Plate-Out in PVC Extrusion. ll. Lubricant
Effects on the Formation of Die Plate-Out in Lead-Based Rigid PVC Formulations,” J. Vinyl
Addit. Technol., vol. 18, no. 2, pp. 209–215, 2012.
[8] J. E. McGrath, “Chain reaction polymerization,” J. Chem. Educ., vol. 58, pp. 844–861, 1981.
[9] D. Braun, H. Cherdron, M. Rehahn, H. Ritter, and B. Voit, Polymer Synthesis: Theory and
Practice. Berlin Heidelberg: Springer, 2013.
[10] P. A. Lovell and M. S. El-Aasser, “Emulsion Polymerization and Emulsion Polymers.” Wiley,
Manchester, Bethlehem, 1997.
[11] B. Brooks, “Suspension polymerization processes,” Chem. Eng. Technol., vol. 33, no. 11, pp.
1737–1744, 2010.
[12] M. J. Rheem, H. Jung, J. U. Ha, S.-H. Baeck, and S. E. Shim, “Suspension polymerization of
thermally expandable microspheres using low-temperature initiators,” Colloid Polym. Sci., vol.
295, no. 1, pp. 171–180, 2017.
[13] T. Furuncuolu, D. Ugur, and V. Aviyente, “Role of Chain Transfer Agents in Free Radical
Polymerization Kinetics,” Macromolecules, vol. 43, pp. 1823–1835, 2010.
[14] C. C. Han, “Molecular weight and temperature dependece of intrisic viscosity of polymer
solutions,” Polymer (Guildf)., vol. 20, pp. 1083–1086, 1979.
[15] M. Raphael and S. Rohani, “On-line estimation of solids concentrations and mean particle size
using a turbidimetry method,” Powder Technol., vol. 89, no. 2, pp. 157–163, 1996.
[16] R. J. M. Tausk, H. C. J. Venselaar, H. C. Corbet, and P. N. Wilson, “Light Transmission
Instrument for Particle Size Analysis of Colloidal Dispersions,” Power Technol., vol. 27, pp.
215–218, 1980.
[17] B. R. Jennings and H. G. Jerrard, “Rayleigh-Gans-Debye and Mie theories in the determination
of spherical particle sizes,” J. Colloid Sci., vol. 20, no. 5, pp. 448–452, 1965.
[18] G. Butters, Particulate Nature of PVC: Formation, Stucture an Processing. London: Applied
Science Publishers, 1982.
[19] A. . Fallis, Principles of Polymer Chemistry, vol. 53, no. 9. Berlin Heidelberg: Springer, 2013.
[20] F. Puoci, Advanced Polymers in Medicine. Berlin Heidelberg: Springer, 2014.
[21] C. S. Chern, “Emulsion polymerization mechanisms and kinetics,” Prog. Polym. Sci., vol. 31, no.

107
5, pp. 443–486, 2006.
[22] H. Tobita, A. E. Hamielec, H. Tobita, and A. E. Hamielec, “Polymerization Processes, 2.
Modeling of Processes and Reactors,” Ullmann’s Encyclopedia of Industrial Chemistry. Wiley,
pp. 1–49, 2015.
[23] H. G. Yuan, G. Kalfas, and W. H. Ray, “Suspension Polymerization,” Polym. Rev., vol. 31, no. 2–
3, pp. 215–299, 1991.
[24] C. Kotoulas and C. Kiparissides, “A generalized population balance model for the prediction of
particle size distribution in suspension polymerization reactors,” Chem. Eng. Sci., vol. 61, no.
2, pp. 332–346, 2006.
[25] E. Vivaldo-lima, P. E. Wood, A. E. Hamielec, A. Penlidis, and P. B. Equation, “An Updated
Review on Suspension Polymerization,” Ind. Eng. Chem. Res., vol. 36, pp. 939–965, 1997.
[26] R. O. Ebewele, Encyclopedia of Polymer Science and Technology. New York: CRC Press, 2000.
[27] E. G. Chatzi and C. Kiparissides, “Dynamic simulation of bimodal drop size distributions in low-
coalescence batch dispersion systems.,” Chem. Eng. Sci., vol. 47, no. 2, pp. 445–456, 1992.
[28] M. Villalobos, Aurelio, Synthesis and modelling of high Tg copolymers through suspension
copolymerization with bifunctional initiators. Ontario: McMaster University, 1992.
[29] I. Fischer, W. F. Schmitt, H.-C. Porth, M. W. Allsopp, and G. Vianello, “Poly (Vinyl Chloride),”
Ullmann’s Encyclopedia of Industrial Chemistry. Wiley, pp. 1–29, 2014.
[30] P. Lorz, F. Towae, W. Enke, R. Jäckh, N. Bhargava, and W. Hillesheim, “Plastics, Additives,”
Ullmann’s Encyclopedia of Industrial Chemistry. Wiley, pp. 620–669, 2012.
[31] C. Wang, H. Wang, J. Fu, and G. Gu, “Effects of additives on PVC plastics surface and the
natural flotability,” Colloids Surfaces A Physicochem. Eng. Asp., vol. 441, pp. 544–548, 2014.
[32] F. Yang and V. Hlavacek, “Improvement of PVC wearability by addition of additives,” Powder
Technol., vol. 103, no. 2, pp. 182–188, 1999.
[33] K. E. Atkins, Plastics Additives - An A-Z reference. Dordrecht: Springer Science + Business
Medial, 1998.
[34] P. Lorz, F. Towae, W. Enke, R. Jäckh, N. Bhargava, and W. Hillesheim, “Plasticizers,” Ullmann’s
Encyclopedia of Industrial Chemistry. Wiley, pp. 599–617, 2012.
[35] P. Lorz, F. Towae, W. Enke, R. Jäckh, N. Bhargava, and W. Hillesheim, “Plastics Processing, 1.
Processing of Thermoplastics,” Ullmann’s Encyclopedia of Industrial Chemistry. Wiley, pp.
155–192, 2012.
[36] K. Van Everbroeck, “PVC training: Fusion-Gelation,” Westerlo-Oevel.
[37] E. E. Lacatus and C. E. Rogers, “The effect of fusion and physical aging on the toughness of
poly(vinyl chloride),” J. Vinyl Technol., vol. 8, no. 4, pp. 183–188, 1986.
[38] V. Chloride and P. Morphology, “Poly ( Vinyl Chloride ) Processing Morphology,” J. vinyl
Technol., vol. 2, no. 3, pp. 165–168, 1980.
[39] H. Zweifel, R. Maier, and M. Schiller, Plastics Additives Handbook. Cincinnati, Ohio: Hanser,
2001.
[40] M. Gilbert, N. Varshney, K. Van Soom, and M. Schiller, “Plate-out in PVC extrusion: I. Analysis
of plate-out,” J. Vinyl Addit. Technol., vol. 14, no. 1, pp. 3–9, 2008.
[41] D. S. van Es et al., “The compatibility of (natural) polyols with heavy metal- and zinc-free
poly(vinyl chloride): Their effect on rheology and implications for plate-out,” Polym. Degrad.

108
Stab., vol. 93, no. 1, pp. 50–58, 2008.
[42] J. P. Disson and S. Girois, “Acrylic process aids for PVC: From theoretical concepts to practical
use,” J. Vinyl Addit. Technol., vol. 9, no. 4, pp. 177–187, 2003.
[43] W. Lau, M.-H. L. Sheng, and S.-Y. Hsu, “Acrylic synthetic lubricant,” 60/739,985, 2005.
[44] M. Robert, “Lubricating composition,” 11169873.4, 2011.
[45] Y. H. Zhang, Y. Wang, X. C. Kong, and D. M. Zhao, “The Polarity of poly(methyl methacrylate)
Copolymers,” Appl. Mech. Mater., vol. 687–691, pp. 4411–4414, 2014.
[46] E. Moghbelli, R. Banyay, and H. J. Sue, “Effect of moisture exposure on scratch resistance of
PMMA,” Tribol. Int., vol. 69, pp. 46–51, 2014.
[47] G. Strobl, Physics of Polymers: Concepts for Understanding Their Structures and Behavior.
Berlin Heidelberg: Springer, 2007.
[48] W. Hu, Polymer Physics: A Molecular Approach. Wien: Springer, 2013.
[49] J. W. Summers, “Lubrication mechanism of poly(vinyl chloride) compounds: Changes upon
PVC fusion (gelation),” J. Vinyl Addit. Technol., vol. 11, no. 2, pp. 57–62, 2005.
[50] S. A. McGlashan, V. T. O’Brien, K. Awati, and M. E. Mackay, Polymer Rheology, vol. 37.
Munich: Hanser, 1998.

109
110
Attachments
Table 37: Reaction conditions and final product properties of all performed polymerization tests

Recipe AcrLub6* AcrLub_9 AcrLub_10A** AcrLub10_B AcrLub11 AcrLub12 AcrLub13 AcrLub14 AcrLub15

Reactor R1 R2 R1 R2 R1 R2 R1 R2 R1 R2 R1 R2 R1 R2 R1 R2

Initiator in P4: 0.75S P4: A81: P4: P4: P4: P4: P4: P4: S P4: S P4: S P4: S P4: S P4: S P4: P4: S P4: S
monomer 0.75S 0.38Q 0.75S 0.75S 0.75S 0.75S 0.75S A81: 0.5Q A81: Q A81: Q 2.00S A81: Q A81: Q
emulsion(PHR) A81: Q
Temp. (°C) 70 70 70 90 90 90 90 90 90

Monomers M/L A/B 0/B A/B A/B A/B A/B A/B A/B A/B
(charge 3a) (PHR)
F (charge 1) (PHR) 0.1 0.1 0.1 0.1 0.4 0.1 0.1 0.1 0.1 0.1 0.1 0.1 0.1 0.1 0.1 0.1 0.1

M1 (charge 3a) 0 0.0 0.0 0.0 0.0 0.0 0.075U 0.15U 0.15U 0.23U 0.23U 0.23U 0.23U 2.5U 2.5U U 1.75U
(PHR)
Emulsifier( charge EM17: 0.3 EM17: EM17: EM17: EM17: EM17: EM17: EM17: EM17: EM17: EM15: EM15 EM15 EM15 EM15 EM15 EM15
1) (PHR) 0.3 0.3 0.3 0.3 0.3 0.3 0.3 0.3 0.3 0.0 :0.0 :0.0 :0.0 :0.0 :0.0 :0.0
Emulsifier in EM17: 2.0 EM17: EM17: EM17: EM17: EM17: EM17: EM17: EM17: EM17: EM15: EM15 EM15 EM15 EM15 EM15 EM15
monomer emulsion 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0 0.6 :0.6 :0.6 :0.6 :0.6 :0.6 :0.6
(charge 3a) (PHR)
Mono dosing time 300 300 300 150 150 110 110 110 110 143 100 95 110 95 110 115 115
(min)
Specific viscosity 10.57X 7.26X 12.63X 10.17X 7.91X 8.09X 3.94X 2.89X 2.51X 2.17X 2.43X 2.20X 2.29X 0.71X 0.77X X 0.83X
(0.4%)
SC (%) / 29.17 15.79 / / 27.59 27.83 28.15 27.91 27.97 28.15 27.66 28.01 28.52 28.96 29.05 28.45

Conv. (PHR/u) / 14.80 16.70 / / 52.68 52.70 / / 42.3 60.8 62.6 55.3 56.8 55.04 53.5 54.2

GPC MW (g/mol) / / / / / 52.15Y / 9.81Y / 10.01Y 6.59Y 5.71Y 5.24Y / / / /

*test result from preliminary research


**the continuous dosing of F failed. During the test it was decided to increasing dosing speed of the monomer emulsion in order to see the influence on final eta-sp.

111
Reactor 1
Reactor 2
Temperature (°C)

Start MM addition
P7: Charge 4
H16 en F: Charge 6
H16: Charge 7
End of MM dosing
F: Charge 8
F: Charge 9

Time (h)

Figure 54: Temperature curve of AcrLub9_1 (orange), AcrLub9_2 (blue) and the bath (yellow) during polymerization. Herein, important points are indicated

112
7,00E+04

NWG w/o E and G11 NWG w/o E and G11 fit

NWG w/o G11 NWG w/o G11 fit

NWG w/o Licowax E NWG w/o Licowax E fit


Shear viscosity (Pa s)

NWG NWG fit

7,00E+03

7,00E+02
1,00E+00 1,00E+01 1,00E+02 1,00E+03

Shear rate (s -1)

-1
Figure 55: Shear viscosity (-) versus shear rate (s ) for the measured and fitted date for the different NWG formulations with adjusted axes

113
Auteursrechtelijke overeenkomst

Ik/wij verlenen het wereldwijde auteursrecht voor de ingediende eindverhandeling:


Development and evaluation of an acrylic lubricant for PVC formulations.

Richting: master in de industriële wetenschappen: chemie


Jaar: 2017

in alle mogelijke mediaformaten, - bestaande en in de toekomst te ontwikkelen - , aan de


Universiteit Hasselt.

Niet tegenstaand deze toekenning van het auteursrecht aan de Universiteit Hasselt
behoud ik als auteur het recht om de eindverhandeling, - in zijn geheel of gedeeltelijk -,
vrij te reproduceren, (her)publiceren of distribueren zonder de toelating te moeten
verkrijgen van de Universiteit Hasselt.

Ik bevestig dat de eindverhandeling mijn origineel werk is, en dat ik het recht heb om de
rechten te verlenen die in deze overeenkomst worden beschreven. Ik verklaar tevens dat
de eindverhandeling, naar mijn weten, het auteursrecht van anderen niet overtreedt.

Ik verklaar tevens dat ik voor het materiaal in de eindverhandeling dat beschermd wordt
door het auteursrecht, de nodige toelatingen heb verkregen zodat ik deze ook aan de
Universiteit Hasselt kan overdragen en dat dit duidelijk in de tekst en inhoud van de
eindverhandeling werd genotificeerd.

Universiteit Hasselt zal mij als auteur(s) van de eindverhandeling identificeren en zal geen
wijzigingen aanbrengen aan de eindverhandeling, uitgezonderd deze toegelaten door deze
overeenkomst.

Voor akkoord,

Coomans, Shana

Datum: 12/06/2017

S-ar putea să vă placă și