akron 1158067153

222
CURE AND MECHANICAL PROPERTIES OF CARBOXYLATED NITRILE RUBBER (XNBR) VULCANIZED BY ALKALINE EARTH METAL COMPOUNDS A Dissertation Presented to The Graduate Faculty of The University of Akron In Partial Fulfillment of the Requirements for the Degree Doctor of Philosophy Tulyapong Tulyapitak December, 2006

Upload: lean-seey-tan

Post on 02-May-2017

244 views

Category:

Documents


16 download

TRANSCRIPT

Page 1: Akron 1158067153

CURE AND MECHANICAL PROPERTIES OF CARBOXYLATED NITRILE

RUBBER (XNBR) VULCANIZED BY ALKALINE EARTH

METAL COMPOUNDS

A Dissertation

Presented to

The Graduate Faculty of The University of Akron

In Partial Fulfillment

of the Requirements for the Degree

Doctor of Philosophy

Tulyapong Tulyapitak

December, 2006

Page 2: Akron 1158067153

ii

CURE AND MECHANICAL PROPERTIES OF CARBOXYLATED NITRILE

RUBBER (XNBR) VULCANIZED BY ALKALINE EARTH

METAL COMPOUNDS

Tulyapong Tulyapitak

Dissertation

Approved: Accepted: ______________________________ ______________________________ Advisor Department Chair Dr. Gary R. Hamed Dr. Mark D. Foster ______________________________ ______________________________ Co-Advisor Dean of the College Dr. Frank N. Kelley Dr. Frank N. Kelley ______________________________ ______________________________ Committee Chair Dean of the Graduate School Dr. Darrell H. Reneker Dr. George R. Newkome _____________________________ ______________________________ Committee Member Date Dr. Alexei P. Sokolov ______________________________ Committee Member Dr. Ali Dhinojwala ______________________________ Committee Member Dr. Avraam I. Isayev

Page 3: Akron 1158067153

iii

ABSTRACT

Compounds of carboxylated nitrile rubber (XNBR) with alkaline metal oxides and

hydroxide were prepared, and their cure and mechanical properties were investigated.

Magnesium oxide (MgO) with different specific surface areas (45, 65, and 140 m2/g) was

used. Increased specific surface area and concentration of MgO resulted in higher cure

rate. Optimum stiffness, tensile strength, and ultimate strain required an equimolar

amount of acidity and MgO. The effect of specific surface area on tensile properties was

not significant. Crosslink density of XNBR-MgO vulcanizates increased with increased

amounts of MgO. ATR-IR spectroscopy showed that neutralization occurs in two steps:

(1) During mixing and storage, MgO reacts with carboxyl groups (RCOOH) to give

RCOOMgOH. (2) Upon curing, these react bimolecularly to form RCOOMgOOCR and

Mg(OH)2. Dynamic mechanical thermal analysis revealed an ionic transition at higher

temperature, in addition to the glass transition. The ionic transition shifts to higher

temperature with increasing MgO concentration. Like MgO-XNBR systems, cure rates of

XNBR-calcium hydroxide (Ca(OH)2) and XNBR-barium oxide (BaO) compounds

increased with increased content of curing agents. Curing by these two agents resulted in

ionic crosslinks. To ensure optimum tensile properties, equimolar amounts of carboxyl

groups and curing agents were required.

Page 4: Akron 1158067153

iv

Dynamic mechanical analysis revealed the ionic transition in these two systems. It shifted

to higher temperature with increased amounts of curing agents. In contrast to MgO,

Ca(OH)2, and BaO, calcium oxide (CaO) gave results similar to those for thermally cured

samples. No ionic transition was observed in XNBR-CaO systems. Tensile strength of

XNBR depended on the strength of ionic crosslinks, which was dependent on the size of

the alkaline metal ions.

Page 5: Akron 1158067153

v

ACKNOWLEDGEMENTS

I really appreciate the support and guidance provided by my advisor, Dr. Gary R.

Hamed. Your enthusiasm, helpful advises, and attention to detail have motivated, and

inspired me to become a better scientist. I would like to place my special thanks to my

co-advisor Dr. Frank N. Kelley, who helps me through a tough situation.

I would like to thank all my committees, Dr. Darrell H. Reneker, Dr. Alexei P.

Sokolov, Dr. Ali Dhinojwala, and Dr. Avraam I. Isayev for useful comments.

I wish many thanks to Dr. Alan N. Gent for his helpful suggestion, and comments.

I would like to extend my sincere thanks to Mr. Robert Seiple, Dr. Critt Olemacher, and

Mr. John Page for your friendly and unconditional help with instrumental analysis. I am

grateful to my group members for their support and friendships.

Most of all I would like to thank my family members, especially my mother, who

has never given up on me. Without you, I will not be a person I am today; Kia, my

beloved wife, who sacrifices her career for taking care of me. I have learned from the first

day we met that you will never leave me behind, and so will I.

Finally, I would like to thank Royal Thai Government for all kinds of support, and

opportunity.

Page 6: Akron 1158067153

vi

TABLE OF CONTENTS

Page

LIST OF TABLES.............................................................................................................. x

LIST OF FIGURES .......................................................................................................... xii

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

II HISTORICAL REVIEW.............................................................................................. 3

2.1 Vulcanization of Carboxylic Rubbers....................................................................... 3

2.1.1 Sulfur and Peroxide Vulcanization .................................................................... 4

2.1.2 Vulcanization via Reactions of Carboxyl Groups ............................................. 4

2.1.3 Cure Behavior of Carboxylic Rubbers............................................................. 10

2.2 Rubber Reinforcement ............................................................................................ 13

2.2.1 Reinforcement by Particulate Fillers ............................................................... 13

2.2.2 Reinforcement by Thermodynamic Phase Separation..................................... 15

2.2.3 Reinforcement by Reaction-Induced Phase Separation................................... 16

2.3 Tensile Strength of Rubbers.................................................................................... 16

2.4 Ionic Aggregation ................................................................................................... 20

2.4.1 Theory .............................................................................................................. 20

2.4.2 Experimental Evidence .................................................................................... 23

Page 7: Akron 1158067153

vii

2.4.3 Ionic Aggregation Models ............................................................................... 25

2.5 Mechanical Properties of Carboxylated Rubbers ................................................... 26

2.5.1 Effect of Carboxyl Content.............................................................................. 28

2.5.2 Influence of Types of Metal Oxides or Salts ................................................... 30

2.5.3 Effect of Metal Oxide Level ............................................................................ 33

2.5.4 Effect of Specific Surface Area ....................................................................... 35

2.5.5 Effect of Filler.................................................................................................. 36

2.5.6 Effect of Plasticizers ........................................................................................ 40

III EXPERIMENTAL..................................................................................................... 42

3.1 Materials ................................................................................................................. 42

3.1.1 Carboxylated Nitrile Rubber (XNBR) ............................................................. 42

3.1.2 Curing Agents .................................................................................................. 42

3.1.3 Solvents............................................................................................................ 43

3.2 Equipments ............................................................................................................. 43

3.3 Compound Preparation ........................................................................................... 43

3.3.1 XNBR-Magnesium Oxide Compounds ........................................................... 43

3.3.2 XNBR-Peroxide Compounds........................................................................... 45

3.3.3 Compounds of XNBR and Other Metal Oxides or Compounds...................... 46

3.4 Cure Behaviors and Molding.................................................................................. 48

3.5 Molding................................................................................................................... 48

3.6 Tensile Testing........................................................................................................ 49

3.7 Crosslink Density Measurements ........................................................................... 50

3.7.1 Near Equilibrium Stress-Strain measurement.................................................. 50

Page 8: Akron 1158067153

viii

3.7.2 Volume Fraction (Vr) of Rubber by Equilibrium Swelling ............................. 51

3.8 Dynamic mechanical properties.............................................................................. 53

3.9 Infrared spectral analysis ........................................................................................ 53

IV RESULTS AND DISCUSSION................................................................................ 54

4.1 Cure Behaviors........................................................................................................ 54

4.1.1 XNBR-MgO Compositions ............................................................................. 54

4.1.2 XNBR-Dicumyl Peroxide Compositions......................................................... 63

4.1.3 XNBR-CaO Compositions............................................................................... 66

4.1.4 XNBR-Ca(OH)2 Compositions........................................................................ 66

4.1.5 XNBR-BaO Compositions............................................................................... 71

4.2 Crosslink Density Measurements ........................................................................... 71

4.2.1 Thermally-Cured XNBR.................................................................................. 71

4.2.2 XNBR-MgO Vulcanizates ............................................................................... 76

4.2.4 XNBR-CaO Vulcanizates ................................................................................ 79

4.2.5 XNBR-Ca(OH)2 Vulcanizates ......................................................................... 81

4.2.6 XNBR-BaO Vulcanizates ................................................................................ 85

4.2.7 Comparison among Metal Compounds ........................................................... 87

4.3 Tensile Properties.................................................................................................... 91

4.3.1 Thermally Cured XNBR.................................................................................. 91

4.3.2 XNBR-MgO Vulcanizates ............................................................................... 94

4.3.3 XNBR-Peroxide Vulcanizates ....................................................................... 100

4.3.4 XNBR-CaO Vulcanizates .............................................................................. 100

4.3.5 XNBR-Ca(OH)2 Vulcanizates ....................................................................... 103

Page 9: Akron 1158067153

ix

4.3.6 XNBR-BaO Vulcanizates .............................................................................. 108

4.3.7 Comparison of Tensile Properties among Metal Compounds ....................... 108

4.3.8 Comparison between Ionic and Covalent Crosslinks .................................... 116

4.4 ATR-IR Spectroscopy........................................................................................... 117

4.4.1 Thermally Cured XNBR................................................................................ 117

4.4.2 XNBR-MgO Compositions ........................................................................... 125

4.4.3 XNBR-CaO Compositions............................................................................. 132

4.4.4 XNBR-Ca(OH)2 Compositions...................................................................... 138

4.4.5 XNBR-BaO Compositions............................................................................. 144

4.4.6 Comparison among Metal Compounds ......................................................... 145

4.5 Dynamic Mechanical Properties ........................................................................... 152

4.5.1 XNBR-MgO Vulcanizates ............................................................................. 152

4.5.2 XNBR-CaO Vulcanizates .............................................................................. 163

4.5.3 XNBR-Ca(OH)2 Vulcanizates ....................................................................... 167

4.5.4 XNBR-BaO Vulcanizates .............................................................................. 172

4.5.5 Comparison among Metal Compounds ......................................................... 177

V CONCLUSIONS..................................................................................................... 179

REFERENCES ............................................................................................................... 181

APPENDICES ................................................................................................................ 190

APPENDIX A CURE PROPERTIES......................................................................... 191

APPENDIX B TENSILE PROPERTIES ................................................................... 193

APPENDIX C MOLECULAR TRANSITION TEMPERATURE ............................ 204

Page 10: Akron 1158067153

x

LIST OF TABLES

Table Page

2.1 Influence of salt formation on tensile properties of butadiene- methacrylic acid copolymer containing carboxyl group of 0.12 ephr (equivalent per a hundred part of rubber)…………………………………

26

2.2 Tensile properties of gum XSBR vulcanized by 10 phr of divalent metal oxides and hydroxides……………………………………………………..

32

2.3 Active and inactive metal compounds……………………………………

32

2.4 Influence of HAF carbon black loading on mechanical properties of ZnO-cured XNBR………………………………………………………………

37

2.5 Effect of silica types on tensile properties of XNBR vulcanizates………..

39

2.6 Effect of clay and calcium carbonate on tensile properties of ZnO-vulcanized XNBR vulcanizates…………………………………………...

40

3.1 Formulations of XNBR-MgO compounds………………………………...

44

3.2 Mixing method……………………………………………………………

45

3.3 XNBR-DCP formulations…………………………………………………

46

3.4 XNBR-CaO compositions…………………………………………………

47

3.5 XNBR-Ca(OH)2 compositions……………………………………………

47

3.6 XNBR-BaO compositions…………………………………………………

47

3.7 Designation and stoichiometric amount of metal oxides or compounds….

47

3.8 Cure times for compositions……………………………………………… 49

Page 11: Akron 1158067153

xi

4.1 Volume fraction of rubber (Vr), sol content, and crosslink density of the raw XNBR cured at 165 oC………………………………………………

74

4.2 Volume fraction (Vr) of rubber, sol content, and crosslink density (ν) of XNBR cured with different magnesium oxides (120 min at 165 oC)……

77

4.3 Volume fraction (Vr) of rubber, sol content, and crosslink density of XNBR cured with dicumyl peroxide 60 min at 165 oC…………………...

79

4.4 Volume fraction (Vr) of rubber, sol content, and crosslink density of XNBR cured with calcium oxide 1000 min at 165 oC……………………

81

4.5 Volume fraction (Vr) of rubber, sol content, and crosslink density of XNBR cured with calcium hydroxide 240 min at 165 oC…………………

83

4.6 Volume fraction (Vr) of rubber, sol content, and crosslink density of XNBR cured with barium oxide 240 min at 165 oC………………………

85

4.7 The effective ionic radii of Mg++, Ca++, and Ba++ ions with various coordination numbers……………………………………………………...

90

4.8 Characteristic group frequencies of the raw XNBR………………………

122

4.9 Characteristic group frequencies of XNBR-MgO compositions……….....

126

4.10 Characteristic group frequencies of XNBR-Ca(OH)2 samples……………

139

4.11 Characteristic group frequencies of XNBR-BaO samples………………...

145

4.12 Molecular transition temperatures of XN-P1.0, and XN-MgA vulcanizates at a frequency of 1.0 Hz……………………………………..

159

4.13 Molecular transition temperatures of XN-P1.0, and XN-CaO vulcanizates at frequency 1.0 Hz…………………………………………

167

4.14 Molecular transition temperatures of XN-P1.0 and XN-Ca(OH)2 vulcanizates at frequency 1.0 Hz…………………………………………

172

4.15 Molecular transition temperatures of XN-P1.0 and XN-CaO vulcanizates at frequency 1.0 Hz…………………………………………

173

Page 12: Akron 1158067153

xii

LIST OF FIGURES

Figure Page 2.1 Cure rheometry of XNBR containing ZnO of different surface area at

twice stoichiometry (S = 35 m2/g, M = 3.5 m2/g, and L = 0.5 m2/g)……...

11

2.2 (a) Cure rheometry, and (b) bin stability of the ZnO-XNBR compounds along with those of the ZnO2-XNBR compound. (KRYNAC PA-50 is a 50/50 masterbatch of medium acrylonitrile NBR and technical grade ZnO2)………………………………………………………………………

12

2.3 Two dimensional schematic of spherical particles of diameter d arranged on a three dimensional square lattice. (s is particle-particle spacing, and t is the thickness of restricted mobility layer of rubber chains)……………………………………………………………………..

14

2.4 Schematic of an ideal rubber network…………………………………….

18

2.5 Tensile strength of gum NR vulcanizates as a function of 1/Mc for various vulcanization systems. ○ accelerated sulfur; × TMT sulfurless; ● peroxide; ∆ high energy radiation……………………………………….

19

2.6 SAXS profiles of a) low density polyethylene, b) a copolymer of ethylene-methacrylic acid, and c) a sodium salt (90% neutralization) of the copolymer……………………………………………………………...

24

2.7 Tensile properties of carboxylic nitrile rubber (0.099 ephr of COOH) cured by various curing systems. A) 0.2 ephr of ZnO, B) sulfur + zinc dimethyldithiocarbamate (0.003 ephr of unavailable zinc), C) 0.2 ephr of zinc + sulfur + zinc dimethyldithiocarbamate (0.003 ephr of unavailable zinc)………………………………………………………………………..

27

2.8 Tensile strength as a function of carboxyl content in butadiene-methacrylic acid copolymers treated with an excess amount of ZnO (twice stoichiometry)……………………………………………………...

28

Page 13: Akron 1158067153

xiii

2.9 Modulus-temperature behavior of butadiene-methacrylic acid copolymers and their lithium salts. (---) RA1 4.7% acid, (− −) RA2 7.7 % acid, (×) RA3 11.6 % acid, (○) RA1 Li 4.7 % salt, (□) RA2 Li 7.7 % salt, (∆) RA3 Li 11.6 % salt……………………………………………………………...

29

2.10 Tensile properties of carboxylic nitrile rubber (butadiene: acrylonitrile: methacrylic acid; 55:35:10) containing 0.1 ephr of carboxyl groups vulcanized by various metal salts and oxides……………………………..

31

2.11 Effect of ZnO level on tensile properties of the carboxylic nitrile rubber (butadiene: acrylonitrile: methacrylic acid; 55:35:10) containing 0.099 ephr of carboxyl content…………………………………………………..

34

2.12 Effect of ZnO levels on temperature-dependent loss tangent (tan δ) of XSBR……………………………………………………………………..

35

2.13 Influence of specific surface areas and levels of ZnO on abrasion resistance of carboxylated nitrile rubber. (□ 3.0 m2/g, + 4.3 m2/g, ◊ 10.0 m2/g)……………………………………………………………………….

36

2.14 Effect of silica loading on storage modulus (E’) and tan δ of ZnO- vulcanized XNBR (Z0 = 0 phr, Z10 = 10 phr, Z20 = 20 phr, Z30 = 30 phr)...

41

4.1 ODR curves of XNBR cured with type A magnesium oxide (specific surface area of 140 m2/g) at 165 oC. (linear scale)………………………..

56

4.2 ODR curves of XNBR cured with type A magnesium oxide (specific surface area of 140 m2/g) at 165 oC. (log-log scale)………………………

57

4.3 ODR curves of XNBR cured with type B magnesium oxide (specific surface area of 65 m2/g) at 165 oC. (linear scale)………………………….

58

4.4 ODR curves of XNBR cured with type B magnesium oxide (specific surface area of 65 m2/g) at 165 oC. (log-log scale)………………………..

59

4.5 ODR curves of XNBR cured with type C magnesium oxide (specific surface area of 45 m2/g) at 165 oC. (linear scale)………………………….

60

4.6 ODR curves of XNBR cured with type C magnesium oxide (specific surface area of 45 m2/g) at 165 oC. (log-log scale)………………………..

61

4.7 ODR curves of the XNBR cured with 2.0x stoichiometric amounts of different magnesium oxides at 165 oC…………………………………….

62

Page 14: Akron 1158067153

xiv

4.8 ODR curves of XNBR cured with dicumyl peroxide at 165 oC…………..

64

4.9 Delta torque, ΔM = MH – ML, as a function of dicumyl peroxide content……………………………………………………………………..

65

4.10 ODR curves of XNBR cured with calcium oxide at 165 oC. (linear scale)………………………………………………………………………

67

4.11 ODR curves of XNBR cured with calcium oxide at 165 oC. (log-log scale)………………………………………………………………………

68

4.12 ODR curves of XNBR cured with calcium hydroxide at 165 oC. (linear scale)………………………………………………………………………

69

4.13 ODR curves of XNBR cured with calcium hydroxide at 165 oC. (log-log scale)………………………………………………………………………

70

4.14 ODR curves of XNBR cured with barium oxide at 165 oC. (linear scale)………………………………………………………………………

72

4.15 ODR curves of XNBR cured with barium oxide at 165 oC. (log-log scale)………………………………………………………………………

73

4.16 Vr and sol content of thermally cured XNBR as a function of cure time…

75

4.17 Vr and sol content of MgO-cured XNBR as a function of MgO concentration………………………………………………………………

78

4.18 Vr and sol content of XNBR-Peroxide vulcanizates as a function of peroxide concentration…………………………………………………….

80

4.19 Vr and sol content of XNBR-CaO vulcanizates as a function of CaO concentration…………………………………………………………

82

4.20 Vr and sol content of XNBR-Ca(OH)2 vulcanizates as a function of Ca(OH)2 concentration……………………………………………………

84

4.21 Vr and sol content of XNBR-BaO vulcanizates as a function of BaO concentration…………………………………………………………

86

4.22 Vr of XNBR vulcanized by various metal compounds as a function of concentration………………………………………………………………

88

Page 15: Akron 1158067153

xv

4.23 Sol content of XNBR vulcanized by various metal compounds as a function of concentration………………………………………………….

89

4.24 Stress-strain curves of thermally cured XNBR……………………………

92

4.25 Tensile properties of thermally cured XNBR as a function of cure time…

93

4.26 Stress-strain curves of XN-MgA vulcanizates (cured 120 min at 165 oC)………………………………………………………………….

95

4.27 Stress-strain curves of XN-MgB vulcanizates (cured 120 min at 165 oC)………………………………………………………………….

96

4.28 Stress-strain curves of XN-MgC vulcanizates (cured 120 min at 165 oC)………………………………………………………………….

97

4.29 Tensile properties of XNBR cured with different magnesium oxides (cured 120 min at 165 oC)…………………………………………………

98

4.30 Stress-strain curves of XN-peroxide vulcanizates (cured 60 min at 165 oC)………………………………………………………………….

101

4.31 Tensile properties of XN-peroxide vulcanizates (cured 60 min at 165 oC)…………………………………………………………………….

102

4.32 Stress-strain curves of XN-Ca vulcanizates (cured 1000 min at 165 oC)…………………………………………………………………….

104

4.33 Tensile properties of XN-Ca vulcanizates (cured 1000 min at 165 oC)......

105

4.34 Stress-strain curves of XN-Ch vulcanizates (cured 240 min at 165 oC)…..

106

4.35 Tensile properties of XN-Ch vulcanizates (cured 240 min at 165 oC)……

107

4.36 Stress-strain curves of XN-Ba vulcanizates (cured 240 min at 165 oC)…..

109

4.37 Tensile properties of XN-Ba vulcanizates (cured 240 min at 165 oC)…….

110

4.38 300% Modulus of the XNBR vulcanized by various metal compounds….

113

4.39 Tensile strength of the XNBR vulcanized by various metal compounds…

114

4.40 Elongation at break of the XNBR vulcanized by various metal compounds………………………………………………………………...

115

Page 16: Akron 1158067153

xvi

4.41 4.41 300% Modulus of the XNBR vulcanized by different curing agents as a function of Vr…………………………………………………

118

4.42 Tensile strength of the XNBR vulcanized by different curing agents as a function of Vr…………………………………………………………

119

4.43 Elongation at break of the XNBR vulcanized by different curing agents as a function of Vr………………………………………………………....

120

4.44 ATR-IR spectra of uncured and thermally cured XNBR in the range 800 to 4000 cm-1…………………………………………………………..

123

4.45 ATR-IR spectra of uncured and thermally cured XNBR in the range 1550 to 1850 cm-1…………………………………………………………

124

4.46 ATR-IR spectra of the uncured neat XNBR and XNBR-MgO compounds in the range 800 to 4000 cm-1………………………………...

128

4.47 ATR-IR spectra of the uncured neat XNBR and XNBR-MgO compounds in the range 1200 to 2000 cm-1…………………………….....

129

4.48 ATR-IR spectra of the neat XNBR and XNBR-MgO vulcanizates in the range 800 to 4000 cm-1 (cured 120 min at 165 oC)…………………

130

4.49 ATR-IR spectra of the neat XNBR and XNBR-MgO vulcanizates in the range 1200 to 2000 cm-1 (cured 120 min at 165 oC)………………..

131

4.50 ATR-IR spectra of the uncured neat XNBR and XNBR-CaO Compounds in the range 800 to 4000 cm-1………………………………..

133

4.51 ATR-IR spectra of the uncured neat XNBR and XNBR-CaO compounds in the range 1550 to 1850 cm-1……………………………….

134

4.52 ATR-IR spectra of the neat XNBR and XNBR-CaO vulcanizates in the range 800 to 4000 cm-1 (cured 1000 min at 165 oC)………………..

135

4.53 ATR-IR spectra of the neat XNBR and XNBR-CaO vulcanizates in the range 1550 to 1850 cm-1 (cured 1000 min at 165 oC)………………

136

4.54 ATR-IR spectra of the uncured neat XNBR and XN-Ca2.0 compounds, and the neat XNBR and XN-Ca2.0 vulcanizates in the range 1550 to 1850 cm-1 (cured 1000 min at 165 oC)……………………………………. 137

4.55 ATR-IR spectra of the uncured neat XNBR and XNBR-Ca(OH)2 compounds in the range 800 to 4000 cm-1………………………………... 140

Page 17: Akron 1158067153

xvii

4.56 ATR-IR spectra of the uncured neat XNBR and XNBR-Ca(OH)2 compounds in the range 1200 to 2000 cm-1……………………………….

141

4.57 ATR-IR spectra of the neat XNBR and XNBR-Ca(OH)2 vulcanizates in the range 800 to 4000 cm-1……………………………………………...

142

4.58 ATR-IR spectra of the neat XNBR and XNBR-Ca(OH)2 vulcanizates in the range 1200 to 2000 cm-1…………………………………………….

143

4.59 ATR-IR spectra of the uncured neat XNBR and XNBR-BaO compounds in the range 800 to 4000 cm-1………………………………...

146

4.60 ATR-IR spectra of the uncured neat XNBR and XNBR-BaO compounds in the range 1200 to 2000 cm-1…………………………….....

147

4.61 ATR-IR spectra of the neat XNBR and XNBR-BaO vulcanizates in the range 800 to 4000 cm-1………………………………………………...

148

4.62 ATR-IR spectra of the neat XNBR and XNBR-BaO vulcanizates in the range 1200 to 2000 cm-1……………………………………………….

149

4.63 ATR-IR spectra of the neat XNBR, XN-MgA2.0, XN-Ca2.0, XN-Ch2.0, and XN-Ba2.0 vulcanizates in the range 1475 to 1850 cm-1…. 151

4.64 Temperature dependence of dynamic storage modulus (E′) of XN-P1.0 and XN-MgA vulcanizates at frequency 1.0 Hz…………………………..

154

4.65 Temperature dependence of dynamic loss modulus (E″) of XN-P1.0 and XN-MgA vulcanizates at frequency 1.0 Hz…………………………..

155

4.66 Schematic drawing of ion hopping mechanisms (opened and closed circles represent ion pairs)………………………………………………...

156

4.67 Temperature dependence of loss tangent (tan δ) of XN-P1.0 and XN-MgA vulcanizates at frequency 1.0 Hz……………………………….

158

4.68 Effect of specific surface area on dynamic storage modulus (E′) of XN-Mg vulcanizates at frequency 1.0 Hz…………………………………

160

4.69 Effect of specific surface area on dynamic loss modulus (E″) of XN-Mg vulcanizates at frequency 1.0 Hz…………………………………

161

4.70 Effect of specific surface area on loss tangent (tan δ) of XN-Mg vulcanizates at frequency 1.0 Hz……………………………..................... 162

Page 18: Akron 1158067153

xviii

4.71 Temperature dependence of dynamic storage modulus (E′) of XN-P1.0 and XN-Ca vulcanizates at frequency 1.0 Hz……………………………..

164

4.72 Temperature dependence of dynamic loss modulus (E″) of XN-P1.0 and XN-Ca vulcanizates at frequency 1.0 Hz……………………………..

165

4.73 Temperature dependence of loss tangent (tan δ) of XN-P1.0 and XN-Ca vulcanizates at frequency 1.0 Hz………………………………….

166

4.74 Temperature dependence of dynamic storage modulus (E′) of XN-P1.0 and XN-Ch vulcanizates at frequency 1.0 Hz……………………………..

169

4.75 Temperature dependence of dynamic loss modulus (E″) of XN-P1.0 and XN-Ch vulcanizates at frequency 1.0 Hz……………………………..

170

4.76 Temperature dependence of loss tangent (tan δ) of XN-P1.0 and XN-Ch vulcanizates at frequency 1.0 Hz………………………………….

171

4.77 Temperature dependence of dynamic storage modulus (E′) of XN-P1.0 and XN-Ba vulcanizates at frequency 1.0 Hz……………………………..

174

4.78 Temperature dependence of dynamic loss modulus (E″) of XN-P1.0 and XN-Ba vulcanizates at frequency 1.0 Hz……………………………..

175

4.79 Temperature dependence of loss tangent (tan δ) of XN-P1.0 and XN-Ba vulcanizates at frequency 1.0 Hz……………………………….....

176

4.80 Temperature dependence of loss tangent (tan δ) of XNBR cured with 2.0x stoichiometry of various metal compounds at frequency 1.0 Hz…….

178

Page 19: Akron 1158067153

1

CHAPTER I

INTRODUCTION

Elastomers are generally characterized by relatively weak interchain interactions

and lack of symmetry or order within molecules. Altering physical characteristics or

designing rubber molecules with specific functions can be made by introducing

functional monomers into conventional rubbers. Incorporation of carboxyl bearing

monomers into polymer chains increases intra- and intermolecular interactions, resulting

in increased tensile strength with inevitably some loss of extension and recovery

properties. Not only are carboxyl groups regarded as polar functional groups, but also

they can be employed to crosslink rubber molecules or attach them to other molecules or

surfaces.

Carboxylated nitrile rubbers (XNBR) are terpolymers of acrylonitrile, butadiene,

and monomers containing carboxyl groups, such as acrylic and methacrylic acids.

Pendant carboxyl groups provide additional curing sites, and make possible using curing

agents that can react with carboxyl groups. XNBRs exhibit self-reinforcement when

vulcanized by divalent metal oxides. This results from ionic crosslinks that aggregate and

form nanometer size domains that phase-separate from the rubber matrix. These domains

are thought to act as multifunctional crosslinks and fillers, thereby producing high

reinforcement. To obtain optimum tensile properties, about twice the stoichiometric

Page 20: Akron 1158067153

2

amount of ZnO is needed. However, recent ATR-IR studies have shown that

neutralization is essentially complete at about the stoichiometric amount of ZnO.

The main purpose of this research was to probe this paradox by studying the

XNBR/MgO systems. The effect of the specific surface area on cure and mechanical

properties was investigated. The systems of XNBR/CaO, XNBR/BaO, and

XNBR/Ca(OH)2 were also studied.

Page 21: Akron 1158067153

3

CHAPTER II

HISTORICAL REVIEW

A copolymer of butadiene and acrylic acid was first recognized in a French patent

awarded to I.G. Farbenindustrie in 1933.1 In 1946, a carboxylic nitrile rubber was first

recorded in a patent.2 The incorporation of carboxyl functional groups into polymer

chains aimed to alter rubber properties. Because of high polarity of carboxyl groups, the

resulting polymers were regarded as polar rubbers. Brown realized the importance of

carboxyl groups as crosslink sites to achieve non-sulfur vulcanizations.3 Carboxylic

nitrile rubbers have been reviewed in greater detail in extensive publications.4-13

2.1 Vulcanization of Carboxylic Rubbers

Unvulcanized raw rubbers are high molecular weight viscoelastic liquids, which

are inelastic, weak, and completely dissolve in solvents. They cannot be useful unless

vulcanized. Vulcanization is a process in which rubber molecules are linked together to

form a three dimensional infinite network, therefore viscoelastic liquids are converted to

viscoelastic solids. Carboxylic rubbers can be vulcanized using curing agents that react

with carboxyl groups, and also by sulfur and peroxide vulcanizations.4-13

Page 22: Akron 1158067153

4

2.1.1 Sulfur and Peroxide Vulcanization

Carboxylic elastomers can be cured using sulfur or peroxide vulcanization recipes

commonly employed in analogous non-carboxylic rubbers.4-8 Carboxyl groups have little

effect on peroxide vulcanization of carboxylic rubbers. A peroxide-cured carboxylic

nitrile rubber containing 40 phr of FEF black had similar properties to those of an

analogous non-carboxylic one.6 Frank, Kraus, and Haefner14 reported that mercaptan-

modified butadiene methacrylic copolymers cured with cumene hydroperoxide have high

bond strength with steel. Unmodified copolymers underwent self-curing due to residual

peroxide left in the polymers.

2.1.2 Vulcanization via Reactions of Carboxyl Groups

a) Anhydride Formation

Carboxylic rubbers can be cured by utilizing reactions of carboxyl groups. Small

amounts of anhydride linkage may be formed via coupling of carboxyl groups (eq. 1)

when heated under rather severe conditions.8

R C

O

C

O

OH2 C

O

OR R H2O (1)

R = polymer chain

+heat

b) Vulcanization by Amines

Diamines, such as ethylene diamine and hexamethylene diamine, have been used

to vulcanize carboxylic elastomers. Crosslink structures range from ionic to covalent

Page 23: Akron 1158067153

5

bonds, depending on heat history. At low heat history, rubber vulcanizates possessed high

tensile strength and compression set. A decrease in tensile strength and compression set

resulted with increasing heat history. This was interpreted as a result from conversion of

ammonium salt crosslinks to amide crosslinks as shown in equation 2.8, 15

R C

O

OH R NH2H2N

R C

O

O R NHH3N

+

+-

R C

O

O R NH3H3N-

C

O

O R2++ -

C

O

R

R C

O

R NHNH C

O

R + 2 H2O

(2)

R = polymer chain

Cooper reported that copolymers of butadiene and acrylic acid when vulcanized with

N,N,N′,N′-tetramethyl ethylenediamine gave vulcanizates with properties similar to those

vulcanized with sulfur.15 Hexamethylenetetramine and hexamethylenediamine were used

to vulcanize carboxylic elastomers prepared from scrap tires, and vulcanizates with high

hardness and low elongation resulted.16

c) Vulcanization by Epoxy Compounds

Carboxylic elastomers can also be cured by epoxy compounds (eq. 3).5, 7 1,2,3,4-

diepoxybutane was found useful for carboxylic polyacrylates. EP201 resin or 3,4-epoxy-

6-methylcyclohexamethyl-3,4-epoxy-6-methylcyclohexane carboxylated proved to be

suitable for Hycar 1072, a carboxylic nitrile rubber.6

Page 24: Akron 1158067153

6

R C

O

OH2 + R CHH2C CH2CH

O O

R C

O

O O C

O

RR CHCH2 CH2CH

OH OH

(3)

R = polymer chain

Mika reported that epoxy resins were highly effective in curing a carboxylic

rubber, Hycar 1571 latex, and that tertiary amines activated the curing, as shown in

equation 4.17

CH2 CH

O

+NR3 R3N CH2 CH

O

R3N CH2 CH

O

R C

O

OH +

R C

O

O CH2 CH

OH

+ NR3

(4)

Chakraborty and De18 found that 7.5 phr of bisphenol A diglycidylether resin

gave a good compromise in processing and technical properties of XNBR containing 40

phr of FEF black. However, plasticizing effect was observed at higher resin content (20

phr).

Page 25: Akron 1158067153

7

d) Vulcanization by Diisocyanate Compounds

Diisocyanate compounds, such as p-tolyl diisocyanate, p-phenylene diisocynate,

and hexamethylene diisocyanate, can be employed to vulcanize carboxylic nitrile

rubbers.6, 8 However, they were difficult to handle due to scorchiness problems. Tensile

properties of vulcanizates were similar to those obtained from sulfur vulcanization

without metal oxide. Carbon dioxide, a by product of reaction (eq. 5), may cause blowing

of vulcanizates.

R C

O

OH2 +

(5)

R = polymer chain

O OC CR NN

R C

O

O C

O

C

O

O C

O

R NHNH R

R C

O

C

O

R NHNH R + 2 CO2

f) Vulcanization by Radiation

Mladenov and coworkers reported vulcanization of a series of carboxylic styrene

butadiene rubbers using gamma radiation.19 Crosslinkages increased linearly with

carboxyl content at small doses. Curing mechanisms were proposed to involve

decarboxylation to form polymeric radicals, which may attack other molecules at double

bonds followed by coupling to form crosslinks, or recombine with other polymeric

radicals (eq. 6).

Page 26: Akron 1158067153

8

R C

O

OHRadiation

R + + HCO2

R + CH

C

O OH

CH2CH2 CH CH CH2 n

CH

C

O OH

CH2CH2 CH CH CH2 nR

CH

C

O OH

CH2CH2 CH CH CH2n

R

Coupling

CH

C

O OH

CH2CH2 CH CH CH2 nR

R R+ R R

(6)

g) Vulcanization by Metal Oxide and Salts

Brown and Duke4 pointed out that carboxylic nitrile rubbers can be vulcanized by

neutralizing carboxyl groups with oxides and salts of polyvalent metals, such as Zn, Pb,

Cd, Mg, and Ca. Vulcanizates with high gum strength can be obtained by using only

ZnO as a curing agent. Tensile properties depend on the levels of ZnO and carboxyl

groups.

Page 27: Akron 1158067153

9

Stoichiometrically, curing reaction of carboxylic rubbers by divalent metal oxide

can be written as equation 7;

R C

O

OH2 MO+ R C

O

O C

O

M O R H2O+ (7)

R = polymer chain

However, Brown and Gibbs5 found that the amounts of zinc bound to carboxyl groups are

chemically equal to the carboxyl content of the polymer. Brown8 suggested that ZnO may

react with carboxyl groups to form the basic salt, – COOZnOH, and also react with

carboxyl groups from the same chain as well as those from different chains.

Monovalent metal can impart a degree of crosslinking. A butadiene methacrylic

copolymer when cured by sodium hydroxide possessed better tensile properties than the

cured neat one.6 Dolgoplosk and coworkers20 obtained similar results on studies of a

terpolymer of butadiene (73.2%), styrene (25.5%), and methacrylic acid (1.5%)

vulcanized by sodium hydroxide.

f) Vulcanization by Combination of Metal Oxides and Sulfur or Peroxide

Generally, sulfur type recipes always contain zinc oxide. In TMTD-accelerated

sulfur vulcanization, reactions between zinc oxide and carboxyl groups are very fast and

dominate at short curing cycles, resulting in a vulcanizate with high tensile strength but

poor compression set and stress relaxation-dependent properties. At longer cure times,

slower sulfur vulcanization takes over, resulting in a vulcanizate with improved

Page 28: Akron 1158067153

10

compression set but lower tensile strength. Similar observations were made in

ZnO/Sulfur/TMTM, ZnO/Sulfur/ZDMC, ZnO/Sulfur/MBTS, and ZnO/Sulfur/CBS

systems.4, 6-8

Chakraborty21 reached the same conclusion for XNBR cured by dual curatives of

sulfur and zinc peroxide. Beekman and Hastbacka22 reported that when half of the ZnO in

ZnO-activated sulfur vulcanization is replaced by magnesium oxide or magnesium

hydroxide, the increased surface activity of MgO resulted in the increased cure rates. The

mixed vulcanization systems of zinc peroxide/ sulfur/ peroxide were studied in XNBR.21

For the XNBR cured by both DCP and metal oxides, peroxide and metal oxide curing

proceeded independently.6, 21

2.1.3 Cure Behavior of Carboxylic Rubbers

Surface area and the amount of zinc oxide play an important role in governing the

cure behavior of carboxylic rubbers vulcanized only by ZnO23, 24 or by both peroxide and

ZnO.25 Cure rate increases with increasing surface area and concentration of ZnO.24

Figure 2.1 shows the role of specific surface area of ZnO on the cure behavior of ZnO-

XNBR compounds. However, carboxylic rubbers vulcanized by zinc oxide suffer from

scorchiness because zinc oxide is very reactive towards carboxyl groups.6, 8, 26

Compounds even cure during milling or storage. Humidity seriously affects Mooney

scorch time of carboxylic and non-carboxylic rubbers.9, 27 Three approaches employed to

solve scorch problems are: i) the use of scorch controllers,6, 8 ii) the use of coated ZnO,28

and iii) the use of zinc peroxide.9, 11, 29

Page 29: Akron 1158067153

11

Figure 2.1 Cure rheometry of XNBR containing ZnO of different surface area at twice stoichiometry (S = 35 m2/g, M = 3.5 m2/g, and L = 0.5 m2/g).24

Organic acids, organic acid anhydrides, silica, boric acid, amines and basic

organic reagents were used as cure retarders and controllers for metal oxide-vulcanized

carboxylic rubbers. Phthalic anhydride, stearic acid, sebacic acid, and succinic anhydride

Page 30: Akron 1158067153

12

were the most effective.6, 8 According to Zakharov and Shadricheva,30 maleic anhydride

was effective for carboxylic styrene butadiene rubbers. These substances not only

reduced scorchiness but also improved tensile and flow properties of compounds.

Hallenbeck28 found that the use of zinc sulfide- and zinc phosphate-coated ZnO instead

of ZnO can improve scorch safety and bin stability of the carboxylated NBR and BR

compounds without affecting final physical properties. The use of metal alkoxides, such

as aluminium isopropoxide, and aluminium ethoxide, along with standard ZnO also gave

a similar effect.28 Zinc peroxide (ZnO2) has been reported to improve scorchiness and

shelf life of carboxylic rubber compounds.9, 11, 27, 29 Cure rheometry and bin stability of

the ZnO-XNBR compound compared to those of the ZnO2-XNBR compound are shown

in Figure 2.2. Zinc peroxide gives much better scorch safety and bin stability than ZnO.

Figure 2.2 (a) Cure rheometry, and (b) bin stability of the ZnO-XNBR compounds along with those of the ZnO2-XNBR compound. (KRYNAC PA-50 is a 50/50 masterbatch of medium acrylonitrile NBR and technical grade ZnO2)11

Page 31: Akron 1158067153

13

2.2 Rubber Reinforcement

Uncrosslinked rubbers are highly entangled molecular chains and viscoelastic.

They can creep and flow under applied forces. They become stiff and more elastic when

chemically crosslinked. However, they have little strength. To be useful, reinforcement is

required. Reinforcement refers to the stiffness and strength imparted to a rubber

vulcanizate by incorporating small hard domains. This can be achieved by many

approaches, for example, by addition of particulate fillers,31, 32 by thermodynamic phase

separation,33 or by reaction-induced phase separation.34, 35

2.2.1 Reinforcement by Particulate Fillers

Reinforcement of rubbers by particulate fillers has been the most popular method

for decades. The most commonly used particulate fillers are carbon black and silica. The

extent of reinforcement depends on many parameters, such as particle size or surface area,

structure, and surface chemistry of the filler particle.36 The key parameter is particle size

or surface area. To give substantial reinforcement, particle size of fillers must be less than

1 μm.32, 36 With increasing surface area (smaller particles), modulus, strength and

abrasion resistance generally increase. Structure, the term used to describe morphology of

fillers, is another important parameter. High structure fillers increase strength and

stiffness.32 Surface chemistry influences physical and chemical interactions between filler

particles and the rubber matrix. Although chemical interactions at the filler-rubber

interface enhance reinforcement, they are not necessary. Physical interactions seem to be

more important.32, 36

Page 32: Akron 1158067153

14

Hamed and Hatfield37 simply modeled how particle size can affect particle-

particle spacing in a particulate-filled rubber. They assumed a volume fraction ν of a

spherical filler of diameter d is dispersed on a three dimensional square lattice of a

continuum rubber matrix (Figure 2.3).

Figure 2.3 Two dimensional schematic of spherical particles of diameter d arranged on a three dimensional square lattice. (s is particle-particle spacing, and t is the thickness of restricted mobility layer of rubber chains)37

According to the model, the nearest neighboring particle spacing is given by:

⎟⎠⎞

⎜⎝⎛ −

ν= 1806.0ds 31 (8)

Page 33: Akron 1158067153

15

which is valid for ν ≤ 0.524. Assuming that each particle is surrounded by a restricted

mobility rubber layer of thickness t, the volume fraction νt of the rubber phase within t

can be calculated by:

ν−⎥⎥⎦

⎢⎢⎣

⎡−⎟

⎠⎞

⎜⎝⎛ +ν

=ν1

1dt21

3

t (9)

Equation 9 is valid for t < s/2, because at t = s/2 the restricted rubber from adjacent

particles begins to overlap. By taking ν = 0.25, and t = 2.8 nm, a micron-sized particle

will result in a small volume of restricted mobility rubber (νt = 0.0056). However, for a

20 nm particle, at the same ν and t, νt = 0.3657. Clearly, the amount of rubber with

restricted mobility is greater in composites containing smaller particle fillers.

Mobility of the rubber phase in a composite with micron-sized particles is very

much like that in bulk unfilled rubber. However, for a composite containing very fine

particles, the rubber matrix may behave differently from unfilled rubber. Restricted

rubber chains increase energy dissipation and may result in crack splitting, which reduces

local stress concentration and inhibits catastrophic growth of the crack.38

2.2.2 Reinforcement by Thermodynamic Phase Separation

Another approach to create hard domains uniformly dispersed throughout the

rubber matrix is by thermodynamic phase separation. Styrenic thermoplastic elastomers,

such as poly (styrene-b-butadiene-b-styrene) or SBS, and poly (styrene-b-isoprene-b-

Page 34: Akron 1158067153

16

styrene) or SIS, form two phases; rigid domains of polystyrene dispersed throughout a

polybutadiene (or polyisoprene) matrix.33, 39 The domain radius of SBS and SIS triblock

copolymers containing polystyrene with molecular weight of about 10,000 g/mole is less

than 20 nm. These rigid domains function both as multiple crosslinks and as filler

particles. An SBS with 27.5 % styrene content was reported to have a tensile strength of

27.1 MPa with elongation at break of 860%,33 comparable to those of a 50 phr N330-

reinforced SBR (23.5 % bound styrene), which has a tensile strength of 28.7 MPa with

ultimate elongation of about 300 %.40

2.2.3 Reinforcement by Reaction-Induced Phase Separation

Substantial reinforcement can also be achieved by blending a rubber with a

compound which can self-react and phase separate to form hard domains. Hydrogenated

acrylonitrile butadiene rubber (HNBR) vulcanized by peroxide and coagent zinc

dimethacrylate (ZDMA) is an example. Upon curing, very fine particles (about 2 nm) of

poly (zinc dimethacrylate) are formed as an in-situ filler, which phase separates from the

HNBR matrix. These primary particles are covalently linked to form secondary ionic

clusters of 20 to 30 nm in size.34, 41 Maximum tensile strength of such a system was

reported to be about 55 MPa with about 500 % ultimate strain.34, 41

2.3 Tensile Strength of Rubbers

Rupture of rubber can occur under a variety of imposed mechanical conditions

such as on stretching to break, during abrasion, or deformations under small cyclic

loading. A corresponding measure of resistance to failure or strength is created for each

Page 35: Akron 1158067153

17

type of rupture. The simplest method is the tensile test, in which the rubber sample is

subjected to a uniform uniaxial tension. Tensile strength and breaking strain are two

important properties used to establish the influence of the nature of rubber and test

conditions.42

When the rubber sample is subjected to simple extension, only a small number of

rubber molecules crossing the fracture plane actually undergo rupture, while most of

rubber molecules remain unaffected. When a crack grows, those molecules will break

successively.43 Consider an ideal network consisting of network chains of molecular

weight Mc between crosslinks arranged in space as shown in Figure 2.4. Upon stretching,

assuming that the load must be carried only by rubber molecules parallel to the direction

of extension, Bueche44 showed that tensile strength (TS) of the ideal network is given by

c

32

c

A FM3NTS ⎟⎟

⎞⎜⎜⎝

⎛ ρ= (10)

where ρ is the density of the rubber, NA is the Avogadro’s number, and Fc is the

maximum load that each molecule can hold. If reasonable values of ρ, Mc and bond

energy (to determine Fc) are taken, the tensile strength calculated from equation 10 is

always greater than the observed value. The deviation arises from neglecting many

important factors, such as, effect of chain ends, distribution of network chain length,

molecular flaws, crystallinity, and viscoelastic effects.44 Many theories45-47 have been

made to explain the tensile strength of rubbers by taking some of these factors into

consideration.

Page 36: Akron 1158067153

18

Figure 2.4 Schematic of an ideal rubber network44

The tensile strength of rubber is influenced by both crosslink types and crosslink

density (Figure 2.5).42 The tensile strength passes through a maximum as the crosslink

density is increased. Flory48 explained that the increased tensile strength of gum NR

vulcanizates with increasing degree of crosslinking before a maximum is attributed to

crystallization of rubber chains upon stretching. At high crosslink density tensile strength

is low because the breaking point is reached before crystallization can occur.

Taylor and Darin46 found similar behavior in gum SBR vulcanizates, which are

not strain-crystallizing. They proposed that chain orientation is a critical factor in

determining tensile strength.

Page 37: Akron 1158067153

19

Figure 2.5 Tensile strength of gum NR vulcanizates as a function of 1/Mc for various vulcanization systems. ○ accelerated sulfur; × TMT sulfurless; ● peroxide; ∆ high energy radiation42

Epstein and Smith49 found that the maximum is greatly dependent on the rate of

extension, and is not shown in swollen samples. Smith and Chu50 studied Viton polymers

and found that the maximum diminishes and finally disappears with increasing

temperature. They concluded that changes in tensile strength with degree of crosslinking

are primarily due to viscoelastic effects, especially energy dissipation.

The tensile strength of rubbers also depends on crosslink structure (Figure 2.5).

Tensile strength decreases in order of increasing strength of crosslink;51

– COO-M+ > – C – S>2 – C – > – C – S2 – C – > – C – S – C – > – C – C –

Page 38: Akron 1158067153

20

Bateman and colleagues51 explained that if crosslinks are weaker than bonds in the main

chain, they will slip and interchange with neighbors. This relieving mechanism will allow

the load to be shared over neighboring chains, and thereby permitting the whole network

to bear higher stress. Tobolsky and Lyons52 studied stress relaxation of rubbers

crosslinked by weak and strong linkages and found no evidence of mechanical lability of

weak crosslinks. They proposed that high tensile strength of rubber crosslinked with

weak bonds is a result of an internally relaxed network, formed at vulcanization

temperatures due to thermal lability of the crosslinks, rather than to relaxation at the

temperature of tensile testing.

2.4 Ionic Aggregation

2.4.1 Theory

The concept of ionic aggregation in metal oxide-vulcanized carboxylic rubbers

was first introduced to explain high tensile strength of vulcanizates by Tobolsky and

coworkers.53 This concept had been proposed by many researchers to account for the

unique behavior of sodium salts of ethylene-methacrylic acid copolymers.54, 55 The first

attempt to treat ionic aggregation theoretically was by Eisenberg.56 He assumed that in a

polymer of low dielectric constant, ionic species would exist fundamentally as contact

ion pairs. This assumption is quite reasonable because the work required to separate ion

pairs is nearly two orders of magnitude greater than the available thermal energy. An

even higher form of ionic aggregates, “multiplet”, would exist in such a system

depending on i) the dimension of the polymer chain and of ion pairs, ii) the tension on the

Page 39: Akron 1158067153

21

chains resulting from ionic aggregation when adjacent ion pairs incorporated into

different multiplets, iii) the electrostatic energy released upon multiplet formation.

The theory assumes that the multiplet is a spherical drop containing only ions and

that polymer chain segments are confined only at the surface of the multiplet. To simplify

calculation, multiplets are assumed to be distributed on a body-centered cubic lattice.

Eisenberg showed that the multiplet radius (rm) is given by

ch

pm S

v3r = (11)

and the number (n0) of ion pairs in the multiplet can be calculated from

ch

m

p

m0 S

Svv

n == (12)

where vp is the volume of an ion pair, vm is the volume of the multiplet, Sm is the surface

area of the multiplet, and Sch is the contact surface of a chain. For a sodium salt of an

ethylene-methacrylic acid copolymer, vp is about 12 Å3 and Sch is about 12 Å2, yielding

rm ~ 3 Å, and vm ~ 100 Å3. For perfect volume occupation, the maximum number of ion

pairs is therefore eight.

Eisenberg also postulated that these multiplets will join together to form larger

aggregates, which he termed “clusters”. Many factors can affect cluster formation, which

are

Page 40: Akron 1158067153

22

i) the work done to stretch the polymer chain upon clustering of multiplets,

ii) the electrostatic energy released on cluster formation, which depends on the

geometry of clustering, and the dielectric constants of the media,

iii) the critical temperature (Tc) at which electrostatic and chain extension

energies are balanced, and

iv) a half of adjacent ion pairs are assumed to be incorporated into the same

cluster.

It is shown that the number of ion pairs per cluster is given by

2332

A

c02

00

c20

2

c

2

c

A

NMn

2r

e4

1K

'kMM

h

hTk3l4

MN

n⎥⎥⎦

⎢⎢⎣

⎡⎟⎟⎠

⎞⎜⎜⎝

⎛ρ

+επ

ρ= (13)

and the distance (R) between clusters is given by

31

A

c

NMn

R ⎟⎟⎠

⎞⎜⎜⎝

⎛ρ

= (14)

where n is the number of ion pairs per cluster, ρ is the density of the polymer, NA is

Avogadro’s number, Mc is the molecular weight of the polymer chain between pendant

ionic groups, l is C – C bond length, k is Boltzmann’s constant, Tc is the critical

temperature, h 2 is the mean square end-to-end distance of the free chain, h 02 is the mean

square end-to-end distance of the freely-jointed chain, M0 is the molecular weight of the

repeat unit, k′ is a parameter related to the particular cluster geometry, K is the dielectric

Page 41: Akron 1158067153

23

constant of the polymer, ε0 is the permittivity of free space, e is the electronic charge, and

r is the distance between the centers of positive and negative charges. Calculations of

intercluster distance were made for a butadiene-sodium methacrylate copolymer

assuming various models. The calculated values are in the range of 44 to 95 Å.

2.4.2 Experimental Evidence

Existence of ionic clusters has been confirmed by many experimental methods,

such as small angle X-ray scattering (SAXS),54, 57, 58 transmission electron microscopy

(TEM),59-61 and dynamic mechanical analysis.59, 62, 63 Figure 2.6 shows the SAXS profiles

of low density polyethylene, a copolymer of ethylene-methacrylic acid, and a sodium salt

(90% neutralization) of the copolymer.54, 57 The peak at low angle (2θ = 4.5o),

corresponding to a spacing of 2 nm, in the profile of the ionomer suggested the presence

of ionic clusters. The peak was observed with all cations, including monovalent and

divalent metals, also ammonium and quaternary ammonium ions.54

Electron microscopy has shown nanometer-sized ionic aggregates.59-61 Marx and

coworkers60 reported ionic aggregates of 1.3 to 2.6 nm in size for butadiene-sodium

methacrylate copolymers. For ZnO-vulcanized XSBR, Sato59 found ionic domains of

about 5 nm uniformly dispersed in the rubber matrix. STEM images of Zn-neutralized

ethylene-methacrylic acid copolymers revealed nearly spherical ionic aggregates of 2.5 to

2.8 nm randomly distributed throughout the polymer matrix.61

Dynamic mechanical studies of ZnO-activated sulfur vulcanization of XSBR

showed, other than the glass transition, a second transition at temperatures 45 to 60 oC in

the temperature-tan δ plot.59 This transition did not appear in the sulfur cured sample

Page 42: Akron 1158067153

24

without ZnO. It is attributed to ionic aggregates. Sato and Blackshaw64 investigated

dynamic mechanical properties of XNBR cured by various metal oxides, and found the

second transition at temperatures 60 to 70 oC.

Figure 2.6 SAXS profiles of a) low density polyethylene, b) a copolymer of ethylene- methacrylic acid, and c) a sodium salt (90% neutralization) of the copolymer54

Page 43: Akron 1158067153

25

Fourier transform infrared spectroscopy (FTIR) has also been employed to study

the morphology of ionomers.65, 66

2.4.3 Ionic Aggregation Models

Many models have been proposed to explain the morphology of ionic aggregates,

such as a hard sphere model,67 a modified hard sphere model,68 and a core-shell model.69

These models are based on interpretation of SAXS profiles of the systems investigated.

Although these models well-explained the SAXS profiles, they were not successful in

describing mechanical properties, especially the appearance of two transitions, the glass

and the ionic transitions, in those ionomers that showed SAXS peaks. The existence of

two transitions indicates that the materials behave like a two-phase system. The

dimensions of the phase-separated region are at least 50 to 100 Å, while the calculated

interspacing between scattering entities from the proposed models is in the order of 30 Å.

This casts a doubt on how to pack 50 to 100 Å particles into a 30 Å lattice.70 The model

that better explains the morphology of ionomers is the Eisenberg-Hird-Moore (EHM)

model.71 This model is based on multiplet formation. The important feature of the model

is that chain mobility in the vicinity of the multiplet is greatly restricted, and the thickness

of the restricted mobility layer is expected not to exceed the persistent length of the

polymer. An individual multiplet effectively acts as a large multifunctional crosslink and

raises the Tg of the polymers, but the restricted layer around the individual multiplet is

not large enough to exhibit its own Tg. The cluster is formed when a number of the

restricted regions overlap in relatively large region (50 to 100 Å), which exhibit its own

Tg, which is significantly higher than that of the unclustered component.

Page 44: Akron 1158067153

26

2.5 Mechanical Properties of Carboxylated Rubbers

Carboxylic rubbers vulcanized by metal oxides or salts exhibit substantial

reinforcement. These vulcanizates have much greater tensile strength and modulus than

those vulcanized by peroxide or sulfur (without ZnO in the recipe), but are poorer in

compression set and properties related to stress relaxation.4-6 Table 2.1 shows influence

of salt formation on tensile properties of copolymer of butadiene and methacrylic acid.6

Table 2.1 Influence of salt formation on tensile properties of butadiene- methacrylic acid copolymer containing carboxyl group of 0.12 ephr (equivalent per a hundred part of rubber).6

Polymer and Treatment Tensile

Strengtha (psi)

Ultimate Elongationa

(%) Raw polymer, 0.12 ephrb of carboxyl group < 100 > 1,600

Treated with 0.12 ephr of aqueous NaOH 1,700 900

Treated with 0.12 ephr of ZnO 6,000 400 Gum sulfur vulcanizate < 500

a Cured 20 min at 132 oC

b ephr = equivalent part per hundred part of rubber

Cooper72-74 proposed that ionic crosslinks interchange under mechanical stress,

and this mechanism will relieve localized stress concentration, resulting in high tensile

strength. Halpin and Bueche75 studied fracture of sulfur- and ZnO-vulcanized carboxylic

nitrile rubbers and suggested that tensile rupture is a viscoelastic effect and unique

properties of ZnO-cured rubber are natural reflections of a sparse crosslink density.

However, according to Tobolsky and coworkers53, the high strength of carboxylic rubbers

Page 45: Akron 1158067153

27

vulcanized by metal oxides resulted from the presence of ionic clusters, which give rise

to a two-phase, reinforced structure.

Although metal oxide crosslinking of carboxylic rubbers enhances tensile strength,

and stiffness, poor compression set and loss of strength at high temperatures are the main

disadvantages. Compromise properties can be achieved by a combination with covalent

crosslink systems; for example, metal oxide combined with sulfur vulcanization (Figure

2.7).6, 8

Figure 2.7 Tensile properties of carboxylic nitrile rubber (0.099 ephr of COOH) cured by various curing systems. A) 0.2 ephr of ZnO, B) sulfur + zinc dimethyldithiocarbamate (0.003 ephr of unavailable zinc), C) 0.2 ephr of zinc + sulfur + zinc dimethyldithiocarbamate (0.003 ephr of unavailable zinc)8

Bhowmick and De76 reported that in XNBR vulcanizates with a mix of sulfur and

metal carboxylate crosslinks, the technical properties are little affected by variations in

sulfur/accelerator ratios. Chakraborty and coworkers77 found that properties of XNBR

vulcanizates formed with mixed sulfur and metal carboxylate crosslinks are guided more

by the ionic crosslinks, especially at long cure times, where they claimed destruction of

Page 46: Akron 1158067153

28

sulfur crosslinks is counterbalanced by formation of ionic crosslinks. In metal oxide-

cured carboxylic rubbers, many important factors can affect the final properties of rubber

vulcanizates, as discussed next.

2.5.1 Effect of Carboxyl Content

For carboxylic rubbers cured with an excess amount of divalent metal oxides, i.e.

twice the stoichiometric amount of ZnO, the tensile strength increases with the increased

carboxyl content as shown in Figure 2.8.8

Figure 2.8 Tensile strength as a function of carboxyl content in butadiene-methacrylic acid copolymers treated with an excess amount of ZnO (twice stoichiometry).8

Page 47: Akron 1158067153

29

Studies by Otocka and Eirich78 have shown that ionic crosslinks enhance the

rubbery modulus in lithium salts of butadiene-methacrylic acid copolymers, and the

degree of enhancement increases with the increased content of carboxylate groups

(Figure 2.9). However, carboxylate links are thermally labile over the entire rubbery zone.

Figure 2.9 Modulus-temperature behavior of butadiene-methacrylic acid copolymers and their lithium salts. (---) RA1 4.7% acid, (− −) RA2 7.7 % acid, (×) RA3 11.6 % acid, (○) RA1 Li 4.7 % salt, (□) RA2 Li 7.7 % salt, (∆) RA3 Li 11.6 % salt

Page 48: Akron 1158067153

30

Ibarra and Alzorriz79, 80 studied ZnO2-XNBR systems and found that crosslink

density and physical properties increase with carboxyl content and curing time.

2.5.2 Influence of Types of Metal Oxides or Salts

Brown6, 8 reported that carboxylic rubbers can be vulcanized using monovalent,

divalent, and multivalent metal compounds. A butadiene-methacrylic acid copolymer

with carboxyl group of 0.12 ephr (equivalent per a hundred part of rubber), when treated

with 0.12 ephr of sodium hydroxide and cured 20 min at 132 oC, showed an improvement

in tensile properties compared to those of the raw polymer (Table 2.1). Theoretical

amounts of sodium carbonate, potassium carbonate, and lithium hydroxide gave similar

results. The terpolymer of 73.2 % butadiene, 25.5 % styrene, and 1.5 % methacrylic acid

treated with sodium hydroxide was reported to have a tensile strength of 6.1 MPa, but it

fell to zero on raising the temperature from 70 to 100 oC.20 Zakharov81 found that rubber

solutions of butadiene-styrene-methacrylic acid terpolymers in isopropylbenzene when

treated with sodium and potassium hydroxides completely gel within 3 and 24 hr,

respectively.

Salts and oxides of multivalent metals can be used to crosslink carboxylic rubbers.

Tensile properties of a carboxylic nitrile rubber vulcanized by various metal oxides and

salts are shown in Figure 2.10.6, 8 ZnO and PbO give vulcanizates with the highest

properties. Dolgoplosk and coworkers82 studied cure and mechanical properties of

carboxylic styrene butadiene rubber (XSBR) vulcanized by oxides and hydroxides of

divalent metals. The results were shown in Table 2.2. Cure rate and mechanical

properties depend considerably on the nature of the metal oxides and hydroxides. The

Page 49: Akron 1158067153

31

best mechanical properties were obtained when cured with magnesium oxide and calcium

hydroxide.

Figure 2.10 Tensile properties of carboxylic nitrile rubber (butadiene: acrylonitrile: methacrylic acid; 55:35:10) containing 0.1 ephr of carboxyl groups vulcanized by various metal salts and oxides

Starmer25 evaluated effectiveness of various metal oxides and hydroxides using a

peroxide curing recipe, and found that these materials fell into two categories, active and

inactive types (Table 2.3). The active materials behaved similarly to ZnO in that they

increased hardness, modulus, and abrasion resistance. The oxides and hydroxides of the

IIA alkaline earth metals, and IIB together with lead appeared to be in this class.

However, calcium oxide, which was expected to be active, gave conflicting results. No

improvement in properties was observed in the case of the inactive ones. The difference

Page 50: Akron 1158067153

32

between the active and inactive materials was that the former had basic groups on the

particle surface, while the latter did not.

Table 2.2 Tensile properties of gum XSBR vulcanized by 10 phr of divalent metal oxides and hydroxides82

Property MgO ZnO CaO PbO CdO Mg(OH)2 Zn(OH)2 Ca(OH)2 Ba(OH)2

Cure time, min 20 10 100 30 120 20 10 80 60

300% Modulus, kg/cm2 44 18 22 30 23 29 29 55 37

Tensile strength, kg/cm2 389 157 132 128 190 220 241 394 249

Relative elongation, % 850 800 760 740 890 835 660 770 675

Residual elongation, % 22 10 22 14 23 15 2 28 18

Table 2.3 Active and inactive metal compounds25

Active Metal Compounds Inactive Metal Compounds

MgO ZnO Mg(OH)2 ZnO2 Ca(OH)2 CdO SrO HgO BaO PbO Ba(OH)2 Pb3O4

MgCO3 SiO2 Al2O3 Al(OH)3 TiO2 ZnS FeO Fe2O3 NiO CuO SnO Sb2O3

Ibarra and Alzorriz83 reported that the cure and tensile properties of CaO-cured

carboxylated nitrile rubbers increase with increased CaO content to reach an optimum,

then dropped with excessive amounts.

Page 51: Akron 1158067153

33

Metal peroxides, such as, zinc peroxide (ZnO2), magnesium peroxide (MgO2),

and calcium peroxide (CaO2), gave vulcanizates with tensile properties comparable to

those of ZnO-cured samples, while scorch safety was much improved. The use of mixed

metal peroxides is also possible.29

Tant and coworkers84 studied the structure and properties of carboxy-terminated

polyisoprene neutralized by various metals. They found that mechanical properties

depend strongly on the neutralizing cations. Elements of groups IA (Na, and K) and IIA

(Mg, Ca, and Ba) formed highly ionic complexes, and the strength of ionic association

increased with decreasing cation size and with increasing cation charge within each group.

2.5.3 Effect of Metal Oxide Level

Mechanical and physical properties of carboxylic rubbers are greatly dependent

on the degree of neutralization. The amount of metal oxides or salts required for complete

neutralization depends upon the carboxyl content of the polymer. For a carboxylic nitrile

rubber, Brown6, 8 found that optimum tensile properties can be achieved by using twice

the stoichiometric amount of ZnO, assuming that each Zn++ ion reacts with two carboxyl

groups. Figure 2.11 shows the dependence of tensile properties of the carboxylic nitrile

rubber on ZnO concentrations. He suggested that in addition to the zinc carboxylate salt,

– COOZnOOC – , the zinc hydroxycarboxylate salt, – COOZnOH, may also form, the

same suggestion made by Dolgoplosk and coworkers.20

Page 52: Akron 1158067153

34

Figure 2.11 Effect of ZnO level on tensile properties of the carboxylic nitrile rubber (butadiene: acrylonitrile: methacrylic acid; 55:35:10) containing 0.099 ephr of carboxyl content.8

Dolkoplosk and coworkers82 obtained similar results in MgO-cured carboxylic

SBR. Sato59 also found the same behavior for carboxylated SBR vulcanized by ZnO.

Based on evidence from dynamic mechanical analysis, which suggested that ionic

crosslinks exist as ionic aggregates, he therefore proposed that the basic salt would

contribute to ionic crosslinks. Furthermore, the increase in the amount of ZnO up to twice

the stoichiometric amount shifted the position of the ionic transition to higher

temperatures (as much as 15 oC), with little change when using greater amounts of ZnO

(Figure 2.12).

Page 53: Akron 1158067153

35

Figure 2.12 Effect of ZnO levels on temperature-dependent loss tangent (tan δ) of XSBR.59

2.5.4 Effect of Specific Surface Area

Starmer25 studied the effect of specific surface area of ZnO on the mechanical

properties of carboxylic nitrile rubbers, and found that Pico abrasion resistance

significantly improves with increasing specific surface areas (Figure 2.13). Beekman and

Hastbacka22 observed similar behavior when replacing a half amount of ZnO with MgO

of different specific surface area. Starmer25 also recognized that specific surface areas

have little effect on tensile strength and ultimate elongation, as did Beekman and

Hastbacka,22 and Hua.23

Page 54: Akron 1158067153

36

Figure 2.13 Influence of specific surface areas and levels of ZnO on abrasion resistance of carboxylated nitrile rubber. (□ 3.0 m2/g, + 4.3 m2/g, ◊ 10.0 m2/g)

2.5.5 Effect of Filler

Reinforcing fillers, such as carbon black and silica are always used to improve the

mechanical properties of non-carboxylic rubbers. In the case of ZnO-vulcanized

carboxylic nitrile elastomers, Brown and Gibbs5 reported that stress-strain properties of

EPC- and whiting-filled rubber vulcanizates are essentially the same. Thus, they

concluded that EPC black does not increase tensile strength, but acts more like a load

extender. Influence of HAF carbon black on the mechanical properties of ZnO-

vulcanized XNBR is shown in Table 2.485 With carbon black, modulus, hardness, and

tear strength increased, but tensile strength changed little.

Page 55: Akron 1158067153

37

Table 2.4 Influence of HAF carbon black loading on mechanical properties of ZnO-cured XNBR85

Ingredients A B C

Krynac 7.50 XNBR Zinc oxide Stearic acid HAF carbon black

100.0 12.0 1.0 0.0

100.0 12.0 1.0 20.0

100.0 12.0 1.0 30.0

Properties Modulus at 100 % elongation (MPa) Modulus at 300 % elongation (MPa) Tensile strength (MPa) Elongation at break (%) Tear strength (kN/m) Hardness (IRHD) Tension set at 100 % elongation (%)

1.8 3.2 33

1150 32 56 9

4.2 11.6 31

1000 45 70 11

6.6 18.0 28 900 48 75 11

Weir and Burkey86 reported that an excellent balance of compound viscosity and

vulcanizate properties, such as hardness, tensile strength, abrasion and flex cut growth

resistance, can be achieved by using semi-reinforcing carbon blacks (N550, N600, and

N774). Types of carbon black had little effect on hardness and tensile strength. Therefore,

they concluded that highly reinforcing carbon blacks are not necessary in the XNBR

formulations. Sato59 reached a similar conclusion in the case of ZnO-cured XSBR filled

with N660. He found that 300 % modulus very much increases, while tensile strength

undergoes little change. Tensile properties strongly depended on the amounts of both zinc

oxide and carbon black.

Shaheen and Grimm87 studied the effect of silica type on the properties of sulfur-

cured XNBR using the recipes shown in Table 2.5, and found that fumed silica gave

vulcanizates with higher modulus, tear strength, and abrasion resistance than those

Page 56: Akron 1158067153

38

obtained from precipitated silica, but poorer compression set. For precipitated silica,

increasing particle size resulted in shorter scorch and cure times, and lower modulus,

tensile strength, and abrasion resistance. Chakraborty and De88 reported that silica and

clay enhance the properties of XNBR vulcanized by a mixed crosslinking system

(sulfur/zinc peroxide); however, silica is more reinforcing. Because carboxyl groups of

polymer chains can react with silanol groups on silica surface, yielding better filler-

polymer interaction, therefore coupling agents were not necessary.

Mandal and Tripathy89 studied the influence of clay and calcium carbonate on the

physical properties of ZnO-cured XNBR (Table 2.6), and found that with increasing filler

contents, modulus, hardness, and tear strength increase at the expense of tensile strength.

In addition, fillers also affect dynamic mechanical properties of XNBR

vulcanizates. Carbon black,59, 62 silica,90 clay, and calcium carbonate89 have been reported

to shift the ionic transition temperature to a higher temperature with increased loadings.

Figure 2.14 shows the effect of silica loading on storage modulus (E′) and tan δ of ZnO-

vulcanized XNBR.90

Page 57: Akron 1158067153

39

Table 2.5 Effect of silica types on tensile properties of XNBR vulcanizates87

Ingredients A B C

Chemigum NX775 XNBR Harwick DSC-18 Stearic acid Dibutylphthalate Wingstay 29 Sulfur, Spider TMTD Pasco 558T ZnO Hi-Sil 233, precipitated silica (0.022 μm) Hi-Sil BP, precipitated silica (0.04 μm) Cab-O-Sil MS-7SD, fumed silica (0.014 μm)

100.0 0.3 2.0 5.0 1.0 0.5 2.0 5.0 30.0

- -

100.0 0.3 2.0 5.0 1.0 0.5 2.0 5.0 -

30.0 -

100.0 0.3 2.0 5.0 1.0 0.5 2.0 5.0 - -

30.0

Cure properties at 163 oC Minimum torque, (N.m) Maximum torque, (N.m) tS2, (min) tc90, (min) tc95, (min) Tensile properties 100 % Modulus, (MPa) 200 % Modulus, (MPa) 300 % Modulus, (MPa) Tensile strength, (MPa) Elongation at break, (%) Hardness, (Shore A) Aging properties; 70 hr at 121 oC in air oven Tensile strength, (MPa) % Change Elongation at break, (%) % Change Hardness, (Shore A) Point change Tear strength, die C, (kN/m) Tear strength, die C, 100 oC, (kN/m) Compression set, 72 hr at 100 oC, (%) Pico abrasion index Pico abrasion index, aged 1 1hr at 149 oC

1.0 8.8 3.2 10.8 16.5

3.7 7.2 11.2 26.6 480 82

23.1 -13 360 -25 87 5

56.0 18.9 41 348 541

0.8 8.9 1.8 4.8 6.5

3.2 5.4 8.5 19.1 490 77

20.0 5

380 -22 84 7

45.5 15.3 25 249 314

1.6 10.4 4.2 12.7 16.5

4.4 7.8 12.5 26.4 490 85

25.9 -2

370 -24 91 6

62.0 21.5 51 417 767

Page 58: Akron 1158067153

40

Table 2.6 Effect of clay and calcium carbonate on tensile properties of ZnO-vulcanized XNBR vulcanizates89

Ingredients A B C D E F G

XNBR ZnO Stearic acid Calcium carbonate Clay

100 12 1 - -

100 12 1 10 -

100 12 1 20 -

100 12 1 30 -

100 12 1 -

10

100 12 1 -

20

100 12 1 -

30 Properties 100 % Modulus, (MPa) 300 % Modulus, (MPa) Tensile strength, (MPa) Elongation at break, (%) Tear strength, (N/cm) Hardness, (IRHD) Tension set, (%)

1.85 3.11 32

1100 32 56 9

1.95 3.95 28

1050 32 63 11

2.20 4.25 27

1020 36 64 11

2.90 5.30 24 900 38 66 12

2.19 3.64 30

1040 32 61 11

2.45 3.95 28

1030 37 63 12

3.11 4.70 27 990 39 65 12

2.5.6 Effect of Plasticizers

Plasticizers greatly affect the mechanical properties of ionomers. Because of this

biphasic nature, the ionic aggregates and hydrocarbon chains can be plasticized

independently by using high and low polarity plasticizers, respectively.91 Dual phase

plasticization is also possible. Makowski and Lundberg92 studied plasticization of metal

sulfonated EPDM with derivatives of stearic acid, and reported that fatty acids, especially

zinc stearate, not only reduced melt rheology of the polymer, but also helped improve the

mechanical properties.

Mandel, Tripathy, and De93 examined the plasticizing effect of ammonia on

properties of gum and filled XNBR vulcanized by ZnO, and found that ammonia

treatment results in a reduction in modulus and tensile strength. Furthermore, a smaller

peak of the ionic transition was observed in ammonia-treated samples. The plasticization

of ZnO-vulcanized XNBR by zinc stearate was also studied.94

Page 59: Akron 1158067153

41

Figure 2.14 Effect of silica loading on storage modulus (E’) and tan δ of ZnO- vulcanized XNBR (Z0 = 0 phr, Z10 = 10 phr, Z20 = 20 phr, Z30 = 30 phr).90

Page 60: Akron 1158067153

42

CHAPTER III

EXPERIMENTAL

3.1 Materials

3.1.1 Carboxylated Nitrile Rubber (XNBR)

Nipol 1072, bound acrylonitrile content (%), 27 ± 1, carboxyl content (ephr), 0.075

± 0.005, Zeon Chemicals L.P..

3.1.2 Curing Agents

a) Dicumyl peroxide: Di-Cup R, GEO Specialty Chemicals.

b) Magnesium oxide:

(i) Elastomag 100, specific surface area of 140 m2/g, Akrochem Corporation.

(ii) Magchem 50, specific surface area of 65 m2/g, Martin Marietta Magnesia

Specialties Inc.

(iii) Magchem 40, specific surface area of 45 m2/g, Martin Marietta Magnesia

Specialties Inc.

c) Calcium oxide, CaO UN1190, Fisher Scientific.

d) Calcium hydroxide, Ca(OH)2 C97-500, Fisher Scientific.

e) Barium oxide, 99.99% Purity, Sigma-Aldrich.

Page 61: Akron 1158067153

43

3.1.3 Solvents

a) Trichloromethane, Reagent grade, EMD Chemicals Inc.

b) Ethyl alcohol, Reagent grade, Fisher Scientific.

3.2 Equipments

a) A 50 cc laboratory internal mixer, Rheocord System 40, a product of Haake

Buchler.

b) Farrel laboratory mill with a roll diameter of 15 cm and length of 30 cm

c) Monsanto ODR R-100

d) Dake press with platen size of 12.5 in x 12.5 in

e) Instron 5567

f) Nicolet 4700 FT-IR spectrometer equipped with SensIR Durascope to utilize

attenuated total reflectance (ATR) measurement

g) DMTA V, Rheometric Scientific

h) Dumbbell die type V according to ASTM D63895

i) Window mold with dimension of 1.0 mm x 93 mm x 118 mm

3.3 Compound Preparation

3.3.1 XNBR-Magnesium Oxide Compounds

Compound formulations are shown in Table 3.1. Three grades of magnesium

oxide (MgO) were used; Elastomag 100 (140 m2/g), Magchem 50 (65 m2/g), and

Magchem 40 (45 m2/g), assigned as “A”, “B”, and “C”, respectively. Each grade of MgO

contained certain amounts of impurities, thus, all recipes were adjusted accounting for the

Page 62: Akron 1158067153

44

impurity content. Numbers in parentheses are the amounts of materials added; the other

numbers are actual phr of MgO. The neat XNBR is designated as XNBR. Compositions

are designated XN-MgL_, where XN represents XNBR, Mg is for MgO, the letter L can

be “A”, “B”, or “C”, indicating a type of MgO, and the suffix is the amount relative to

stoichiometry, assuming that each Mg++ ion can neutralize two carboxyl groups. The

stoichiometric amount of MgO is 1.5 phr.

Table 3.1 Formulations of XNBR-MgO compounds

Ingredients XNBR XN-MgA0.5

XN-MgA1.0

XN-MgA1.5

XN-MgA2.0

XN-MgA3.0

XN-MgA4.0

XN-MgA5.0

Nipol 1072 (XNBR ) 100 100 100 100 100 100 100 100

Elastomag 100 (98.0%)

0.0 (0.0)

0.75 (0.77)

1.5 (1.53)

2.25 (2.30)

3.0 (3.06)

4.5 (4.59)

6.0 (6.12)

7.5 (7.65)

Ingredients XNBR XN-MgB0.5

XN-MgB1.0

XN-MgB1.5

XN-MgB2.0

XN-MgB3.0

XN-MgB4.0

XN-MgB5.0

Nipol 1072 (XNBR ) 100 100 100 100 100 100 100 100

Magchem 50 (98.0%)

0.0 (0.0)

0.75 (0.77)

1.5 (1.53)

2.25 (2.30)

3.0 (3.06)

4.5 (4.59)

6.0 (6.12)

7.5 (7.65)

Ingredients XNBR XN-MgC0.5

XN-MgC1.0

XN-MgC1.5

XN-MgC2.0

XN-MgC3.0

XN-MgC4.0

XN-MgC5.0

Nipol 1072 (XNBR ) 100 100 100 100 100 100 100 100

Magchem 40 (98.0%)

0.0 (0.0)

0.75 (0.77)

1.5 (1.53)

2.25 (2.30)

3.0 (3.06)

4.5 (4.59)

6.0 (6.12)

7.5 (7.65)

Page 63: Akron 1158067153

45

Compounds were prepared (Table 3.2) in a 50 cc internal mixer (Rheocord 40,

Haake Buchler) using a fill factor of 0.85, and rotor speed of 40 rpm. Final mix

temperature was 85 oC to 95 oC. Compounds were milled and sheeted to about 1 mm

thick on an open mill (friction ratio of 1:1.21), and stored in sealed plastic bags in the

dark until further used.

Table 3.2 Mixing method

Time (min) Procedure

0 - 1 Add rubber

1 – 3 Masticate rubber

3 - 4.5 Add MgO

4.5 – 7 Mix rubber with MgO

7 Dump

3.3.2 XNBR-Peroxide Compounds

XNBR-Peroxide compounds were prepared to compare their properties to those of

metal oxide-cured compounds. Peroxide curing gives covalent crosslinks, while metal

oxide curing gives ionic linkages. Formulations of XNBR-peroxide are shown in Table

3.3. Compositions are named as followed; XN-P_, where letters XN are for XNBR, P is

for dicumyl peroxide (DCP), and the suffix indicates phr of DCP used. All compounds

were prepared in the same way as XNBR-MgO compositions, both mixing (Table 3.2)

and milling methods.

Page 64: Akron 1158067153

46

Table 3.3 XNBR-DCP formulations

Ingredients XNBR XN-P0.25

XN-P0.50

XN-P0.75

XN-P1.0

XN-P1.5

XN-P2.0

XN-P3.0

Nipol 1072 (XNBR )

100 100 100 100 100 100 100 100

Dicumyl peroxide - 0.25 0.50 0.75 1.0 1.5 2.0 3.0

3.3.3 Compounds of XNBR and Other Metal Oxides or Compounds

In addition to MgO, calcium oxide (CaO), calcium hydroxide (Ca(OH)2), and

barium oxide (BaO) were also used as curing agents for the XNBR. All these are

compounds of alkaline earth metals (group IIA in periodic table), with a valency of 2.

Compositions of XNBR-CaO, XNBR-Ca(OH)2, and XNBR-BaO are shown in Tables 3.4,

3.5, and 3.6, respectively. In each case, two levels of metal oxides or compounds are

shown. Numbers in parentheses are added amounts; others are actual content, taking into

account purity. Compositions are designated as XN-Aa_, where XN stands for XNBR,

Aa indicates a type (Table 3.7) of metal oxides or compounds, and the suffix is the

amount relative to stoichiometry, assuming that one mole of metal ion (M++) reacts with

two moles of COOH groups. The stoichiometric amount of each metal compound is

shown in Table 3.7. All compositions were mixed and milled using the same procedure

used in preparing XNBR-MgO and XNBR-peroxide compounds.

Page 65: Akron 1158067153

47

Table 3.4 XNBR-CaO compositions

Ingredients XNBR XN-Ca0.5

XN-Ca1.0

XN-Ca1.5

XN-Ca2.0

XN-Ca3.0

XN-Ca4.0

XN-Ca5.0

Nipol 1072 (XNBR ) 100 100 100 100 100 100 100 100

CaO UN1910 (98.0 %) - 1.05

(1.07) 2.10

(2.14) 3.15

(3.21) 4.20

(4.29) 6.30

(6.43) 8.40

(8.57) 10.5

(10.7)

Table 3.5 XNBR-Ca(OH)2 compositions

Ingredients XNBR XN-Ch0.5

XN-Ch1.0

XN-Ch1.5

XN-Ch2.0

XN-Ch3.0

XN-Ch4.0

XN-Ch5.0

Nipol 1072 (XNBR ) 100 100 100 100 100 100 100 100

Ca(OH)2 (97.0 %) - 1.39

(1.43) 2.78

(2.87) 4.17

(4.30) 5.56

(5.73) 8.34

(8.60) 11.1

(11.5) 13.9

(14.3)

Table 3.6 XNBR-BaO compositions

Ingredients XNBR XN-Ba0.5

XN-Ba1.0

XN-Ba1.5

XN-Ba2.0

XN-Ba3.0

XN-Ba4.0

XN-Ba5.0

Nipol 1072 (XNBR ) 100 100 100 100 100 100 100 100

BaO (99.99%) - 2.88 5.75 8.63 11.5 17.3 23.0 28.8

Table 3.7 Designation and stoichiometric amount of metal oxides or compounds

Aa Type of Metal Oxides or Compounds

Stoichiometric Amount (phr)

Mg MgO 1.50 Ca CaO 2.10 Ch Ca(OH)2 2.87 Ba BaO 5.75

Page 66: Akron 1158067153

48

3.4 Cure Behaviors and Molding

Cure behaviors of all compounds were determined according to ASTM D208496

using Monsanto ODR R100. A specimen of each composition weighing 8 to 9 g was put

into an electrically heated chamber, maintained at 165 oC (329 oF). When the chamber

was closed, the internal biconical disk rotor was oscillated at 3o arc within the rubber

sample. A rheometry curve was recorded and cure parameters such as minimum torque

(ML), maximum torque (MH), and scorch time (ts2) were determined from the curve. MH

is the highest torque attained at the specified time when no plateau or maximum torque

was obtained, ts2 is the time at which the torque rises above ML by 2.0 dN.m.

3.5 Molding

Tensile sheets were prepared by compression molding in a window mold (1 mm x

93 mm x 118 mm) at 165 oC. A rectangular sheet (13 to 14 g) of a compound was placed

between two Mylar sheets, and was placed into the mold; Teflon sheets were used if

rubber compounds stuck to Mylar sheets. The mold was placed between the upper and

lower platens of the press. The applied pressure was 25 tons. Cure times of compositions

are specified in Table 3.8. In the case of thermally cured of XNBR, cure times were 60,

120, 240, 500, and 1,000 minutes. After being removed from the press, the mold was

allowed to cool down at room temperature for 15 to 20 min. The tensile sheet was then

taken out of the mold and kept in a sealed plastic bag for 24 to 48 hr before tensile testing.

Page 67: Akron 1158067153

49

Table 3.8 Cure times for compositions

Compounds Cure Time (min)

XNBR-MgO 120 XNBR-Peroxide 60 XNBR-CaO 1,000 XNBR-Ca(OH)2 240 XNBR-BaO 240

3.6 Tensile Testing

Tensile properties at room temperature (25 ± 2 oC) of all vulcanizates were

determined using the Instron machine model 5567 equipped with an extensometer.

Dumbbell specimens were prepared using a type V die according to ASTM D638.95 The

distance between upper and lower clamps was set at 40 mm and the crosshead speed was

100 mm/min, producing a strain rate of 2.5 min-1 (0.042 s-1). The thickness of each

specimen was taken as an average value of three different positions on the narrow section

of the specimen, which was 3.18 mm wide. Two marks 10.0 mm apart were placed on the

narrow section. When the specimen was stretched, the change in the separation of the two

marks was followed by the extensometer. The tensile properties of the XNBR

vulcanizates, such as stress at specified strain, elongation at break, and tensile strength

were calculated from the measured quantities.

Page 68: Akron 1158067153

50

3.7 Crosslink Density Measurements

3.7.1 Near Equilibrium Stress-Strain measurement

Crosslink densities of vulcanizates were determined by equilibrium stress-strain

measurements. A dumbbell specimen about 1.0 mm thick, with two marks 10.0 mm apart,

was clamped at both ends using metal clips. One end was fastened to a steel bar, and the

other end was hung from a small weight. After 30 minutes, the extension was measured

using a cathetometer. Then an additional weight was added, and again the sample was left

for 30 minutes before the extended length was measured. These processes were repeated

until the extension ratio (λ) of the sample was greater than 3.5. The crosslink density of

the specimen was then determined by using the Mooney-Rivlin equation:97-99

λ+=

⎥⎦⎤

⎢⎣⎡

λ−λ

σ 21

2

C2C2

1 (15)

where λ is the extension ratio, σ is the engineering stress, and C1 and C2 are constants. By

plotting the quantity in the left side of equation 15 versus 1/λ, the intercept 2C1 is

obtained. It is related to crosslink density (ν) through equation 16:

RTC2 1=ν (16)

where R is the gas constant (8.314 J/mol.K), and T is the absolute temperature.

Page 69: Akron 1158067153

51

3.7.2 Volume Fraction (Vr) of Rubber by Equilibrium Swelling

Generally, crosslink density of rubber vulcanizates can also be determined by

equilibrium swelling using the well-known Flory-Rehner equation:100

⎟⎠⎞

⎜⎝⎛ −

χ++−⎟⎟⎠

⎞⎜⎜⎝

⎛−=ν

2VV

]VV)V1(ln[V21

r31r

2rrr

s (17)

where ν is the number of moles of tetrafunctional crosslinks per unit volume, Vr is the

volume fraction of rubber in the swollen gel, Vs is the molar volume of the solvent, and χ

is the rubber-solvent interaction parameter. However, ionic crosslinks exist as ionic

aggregates; therefore, it is better to use Vr as a meassure of crosslink density for the

XNBR vulcanizates.

Chloroform was used as a solvent, because its solubility parameter (δ = 9.30) is

similar to that of 75/25 butadiene/acrylonitrile copolymer (δ = 9.38).101

Specimens with dimensions about 1 mm x 2 mm x 25 mm were cut from cured

sheets and weighed on an analytical balance with an accuracy of 0.01 mg (Mi). These

specimens were put into vials (25 mm in diameter and 95 mm in length), and allowed to

swell in 30.0 mL of chloroform for 7 days in the dark. Then, they were taken out of the

solvent, blotted on paper and quickly weighed (Mgel). After drying for 24 hr at room

temperature, the specimens were dried at 70 oC in a vacuum oven for 24 hr. The dry

weight (Mdry) was measured. The volume fraction of rubber can be determined from

Page 70: Akron 1158067153

52

chloroformrubber

rubberr VV

VV

+= (18)

Vchloroform and Vrubber can be calculated using equations 19 and 20, respectively.

chloroform

igelchloroform

MMV

ρ

−= (19)

⎟⎟⎠

⎞⎜⎜⎝

⎛ρ

−ρ

=−=MO

MOi

dry

dryMOdryrubber

fM

MVVV (20)

VMO is the volume of the metal compound. ρdry is the density of the dry rubber compound.

fMO is the weight fraction of the metal compound. ρMO is the density of the metal

compound; the densities of MgO, CaO, Ca(OH)2, and BaO are 3.20, 3.30, 2.24, and 5.75

g/cm3, respectively. Chloroform has a density of 1.473 g/cm3 at 25 oC, and the neat

XNBR has a density of 0.98 g/cm3.

To measure the density of the dry rubber (ρdry), a specimen with dimension 1 mm

x 3 mm x 25 mm was cut from a cured sheet, and its weight in air (Mair) and ethanol

(Mliq) was determined. The weight when immersed in ethanol (ρ =0.785 g/cm3) is less

than that in air by the weight of ethanol displaced. The volume of ethanol displaced is

equal to that of the specimen. The density of dry rubber compound can be calculated

from

liqair

airliqdry MM

M−

ρ=ρ (21)

Page 71: Akron 1158067153

53

3.8 Dynamic mechanical properties

Dynamic mechanical properties of all XNBR vulcanizates were determined using

the Rheometric Scientific DMTA V. The tension mode and strain amplitude of 0.05%

were employed. A frequency of 1.0 Hz was used. A rectangular specimen (25 mm x 6.35

mm x 1.0 mm) was cut from a cured sheet, and then put into a temperature-controlled

chamber. The specimen was cooled down to -80 oC and then held between two clamps

with a gap between them of 10.0 mm. The sample was maintained at -80 oC for 30 min,

and heated from -80 oC to 180 oC at a rate of 2 oC/min. Dynamic mechanical properties,

such as E’, E” and tan δ, were recorded by computer.

3.9 Infrared spectral analysis

ATR-FTIR spectroscopy was employed to study the neutralization of XNBR

compounds. Two FT-IR spectrophotometers, Nicolet 4700 FT-IR and Nicolet 5SXC FT-

IR, were employed. They were equipped with a SensIR Durascope utilizing attenuated

total reflectance (ATR). The former was used to study XNBR-MgO, and XNBR-peroxide

systems; the latter was employed for the rest of the compounds. A specimen about 1.0

mm thick was scanned with a resolution of 4 cm-1 and 215 times, and the final result was

the average of 215 spectra. The cured samples were kept about 2 weeks in a dark

container at room temperature before testing.

For uncured compounds (aged about 2 weeks), specimens were compression-

molded to a thickness of about 1.0 mm.

Page 72: Akron 1158067153

54

CHAPTER IV

RESULTS AND DISCUSSION

4.1 Cure Behaviors

4.1.1 XNBR-MgO Compositions

Cure curves at 165 oC of the neat XNBR and compounds with different

magnesium oxides are shown in Figures 4.1 to 4.6. The neat XNBR stiffens slightly upon

heating, possibly due to self-coupling of carboxyl groups to form anhydride bridges (eq.

1). Cure rheometry of XN-MgA compounds is shown in Figure 4.1. The same results on

log-log scales are given in Figure 4.2, which shows that rheometric torques at early stage

(less than 10 min) of all compounds, except for XN-MgA0.5, are higher than that of the

neat XNBR. Maximum torque of XN-MgA0.5 is slightly lower than that of the neat

XNBR. With the increased amounts of MgO, cure rate increases, while scorch time

decreases. Maximum torques enormously increase with increasing MgO concentrations

up to twice stoichiometry, and little increase thereafter with excess amounts. Therefore,

optimum cure requires at least 2.0x stoichiometric amounts.

ODR curves of the neat XNBR and XN-MgB compounds are given in Figures 4.3

(linear scale), and 4.4 (log-log scale). Similar to XN-MgA compounds, cure rates

increase with the increased MgO contents, while scorch time decreases. However, XN-

MgB compounds cured slower than did XN-MgA compounds. At an early stage of

Page 73: Akron 1158067153

55

heating, compositions containing up to 2.0x stoichiometry have lower torques than does

the neat XNBR. The torque of XN-MgB0.5 remains lower than that of the neat XNBR

for the entire heating time.

Cure rheometry of XN-MgC compounds is shown in Figure 4.5. Torque of XN-

MgC0.5 is lower than that of the neat XNBR for the whole heating time. With increasing

MgO concentration, cure rate and maximum torque increase, while scorch time decreases.

However, cure rates of XN-MgC compounds are lower than those of XN-MgB, and XN-

MgA compositions. The log-log plot of cure rheometry of XN-MgC compounds (Figure

4.6) shows that at early stages of heating the torque of the neat XNBR is higher than that

of XN-MgC compounds containing MgO up to 3.0x stoichiometry, except for XN-

MgC4.0 and XN-MgC5.0. Thermal crosslinking of the XNBR is retarded by the presence

of large surface area MgO. Apparently, some carboxyl groups have reacted with MgO.

Based on ATR-IR results which will be discussed later, neutralization appears to be

involved in two steps, equations 22 and 23, respectively.102 MgO first reacts with

carboxyl groups, resulting in the magnesium hydroxycarboxylate salt, (– COOMgOH),

followed by coupling of hydroxycarboxylate salts to form the magnesium carboxylate

salt, (– COOMgOOC –). The overall reaction is in accord with equation 24, indicating

that equimolar amounts of acidity and MgO are required. The first neutralization does not

yield an elastically effective rubber network, but it decreases the concentration of

carboxyl groups available for thermal crosslinking. At longer time, torques of compounds

rise due to the increase in intermolecular links produced by the second neutralization, or

aggregation of salt products from both neutralization steps.

Page 74: Akron 1158067153

56

0 50 100 150 200 2500

10

20

30

40

50

60

70

80

XN-MgA4.0 XN-MgA5.0

XN-MgA3.0

XN-MgA2.0

XN-MgA1.5

XN-MgA1.0

XN-MgA0.5

XNBR

Tor

que

(dN

.m)

Time (min)

Testing temperature: 165 oC

Figure 4.1 ODR curves of XNBR cured with type A magnesium oxide (specific surface area of 140 m2/g) at 165 oC. (linear scale)

Page 75: Akron 1158067153

57

1 10 1005

6789

1010

20

30

40

50

60708090

100100Testing temperature: 165 oC

XN-MgA5.0XN-MgA4.0

XN-MgA3.0

XN-MgA2.0

XN-MgA1.5

XN-MgA1.0

XN-MgA0.5

XNBRTor

que

(dN

.m)

Time (min)

Figure 4.2 ODR curves of XNBR cured with type A magnesium oxide (specific surface area of 140 m2/g) at 165 oC. (log-log scale)

Page 76: Akron 1158067153

58

0 50 100 150 200 2500

10

20

30

40

50

60

70XN-MgB5.0

XN-MgB4.0

XN-MgB3.0

XN-MgB2.0

XN-MgB1.5

XN-MgB1.0

XN-MgB0.5

XNBR

Tor

que

(dN

.m)

Time (min)

Testing temperature: 165 oC

Figure 4.3 ODR curves of XNBR cured with type B magnesium oxide (specific surface area of 65 m2/g) at 165 oC. (linear scale)

Page 77: Akron 1158067153

59

1 10 1005

6789

1010

20

30

40

50

60708090

100100Testing temperature: 165 oC

XN-MgB3.0

XN-MgB5.0

XN-MgB4.0

XN-MgB2.0

XN-MgB1.5

XN-MgB1.0

XNBR

XN-MgB0.5

Tor

que

(dN

.m)

Time (min)

Figure 4.4 ODR curves of XNBR cured with type B magnesium oxide (specific surface area of 65 m2/g) at 165 oC. (log-log scale)

Page 78: Akron 1158067153

60

0 50 100 150 200 2500

10

20

30

40

50

60

70XN-MgC5.0

XN-MgC4.0

XN-MgC3.0

XN-MgC2.0

XN-MgC1.5

XN-MgC1.0

XNBR

XN-MgC0.5

Tor

que

(dN

.m)

Time (min)

Testing temperature: 165 oC

Figure 4.5 ODR curves of XNBR cured with type C magnesium oxide (specific surface area of 45 m2/g) at 165 oC. (linear scale)

Page 79: Akron 1158067153

61

1 10 1005

6789

1010

20

30

40

50

60708090

100100Testing temperature: 165 oC

XN-MgC5.0

XN-MgC4.0

XN-MgC3.0

XN-MgC2.0

XN-MgC1.5

XN-MgC1.0XNBR

XN-MgC0.5

Tor

que

(dN

.m)

Time (min)

Figure 4.6 ODR curves of XNBR cured with type C magnesium oxide (specific surface area of 45 m2/g) at 165 oC. (log-log scale)

Page 80: Akron 1158067153

62

1 10 1005

6

789

10

20

30

40

50

60

70

XNBR

XN-MgC2.0

XN-MgB2.0

Tor

que

(dN

.m)

Time (min)

XN-MgA2.0

Testing temperature 165 oCA = 140 m2/gB = 65 m2/gC = 45 m2/g

Figure 4.7 ODR curves of the XNBR cured with 2.0x stoichiometric amounts of different magnesium oxides at 165 oC.

Page 81: Akron 1158067153

63

RCOOH + MgO RCOOMgOH (22)

2 RCOOMgOH RCOOMgOOCR + Mg(OH)2 (23)

2 RCOOH + 2 MgO RCOOMgOOCR + Mg(OH)2 (24)

R = polymer chain

Evidently, cure reactions of MgO-vulcanized XNBR are dependent on both the

specific surface area and the concentration of MgO. Compounds cure quickly with

increasing specific surface area and concentration. Figure 4.7 compares XNBR

compositions containing different magnesium oxides at twice stoichiometry. Clearly,

large surface area MgO results in faster curing. Dependence of cure reaction on both

surface area and concentration indicates that it is a diffusion-controlled reaction.23, 24

4.1.2 XNBR-Dicumyl Peroxide Compositions

ODR curves of XN-P compounds are given in Figure 4.8. Minimum torque is

little affected by increased amounts of peroxide. With increasing peroxide content, cure

rates and maximum torques increase, while scorch time decreases. The increase in

maximum torque or stiffness is due to increased crosslink density. Figure 4.9 shows the

dependence of ΔM, MH – ML, on peroxide content.

Page 82: Akron 1158067153

64

0 10 20 30 40 50 600

20

40

60

80

100

Testing temperature: 165 oC

XN-P3.0

XN-P2.0

XN-P1.5

XN-P1.0

XN-P0.75

XN-P0.50

XN-P0.25

Tor

que

(dN

.m)

Time (min)

XNBR

Figure 4.8 ODR curves of XNBR cured with dicumyl peroxide at 165 oC

Page 83: Akron 1158067153

65

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.50

20

40

60

80

100

ΔΜ

= M

H -

ML (d

N.m

)

Amount of peroxide (phr)

Extrapolated line from lower concentration of peroxide

Experiment

Figure 4.9 Delta torque, ΔM = MH – ML, as a function of dicumyl peroxide content.

Page 84: Akron 1158067153

66

The increase in ΔM suggests that crosslink density increases with increasing peroxide

concentration. However, at higher concentrations the slope declines suggesting that the

efficiency of peroxide crosslinking decreases. This may be due to some acidic nature of

the rubber matrix. Acids induce heterolytic or ionic decomposition of the peroxide, in

which the peroxide is consumed without radical formation.103, 104

4.1.3 XNBR-CaO Compositions

Cure rheometry of XNBR cured with calcium oxide is shown in Figure 4.10. The

same results using log-log scales are given in Figure 4.11. The minimum torque is highest

for the neat XNBR. This suggests solubilization and plasticization by calcium oxide.

Thermal crosslinking is initially retarded by presence of CaO particles. Maximum torque

of the raw XNBR is comparable to those of XNBR-CaO compounds, suggesting that

similar levels of cure are reached

4.1.4 XNBR-Ca(OH)2 Compositions

Cure curves of XNBR-Ca(OH)2 compounds are shown in Figure 4.12. A log-log

plot of the same results is given in Figure 4.13. At early stages of heating, torque of the

raw XNBR is higher than those of XNBR-Ca(OH)2 compounds, except for XN-Ch5.0.

Similar to XNBR-MgO and XNBR-CaO systems, thermal crosslinking is retarded by

presence of Ca(OH)2. Torque of XN-Ch0.5 remains lower than that of the neat XNBR for

the entire heating time. Cure behaviors of XNBR-Ca(OH)2 compounds are quite different

from those of XNBR-CaO systems.

Page 85: Akron 1158067153

67

0 200 400 600 800 10005

10

15

20

25

30

35

40

45

Tor

que

(min

)

Time (min)

ODR curves of XN-CaO compounds at 165 oC

XNBR

XN-Ca3.0

XN-Ca4.0

XN-Ca5.0

XN-Ca0.5

XN-Ca1.0

XN-Ca1.5

XN-Ca2.0

Figure 4.10 ODR curves of XNBR cured with calcium oxide at 165 oC. (linear scale)

Page 86: Akron 1158067153

68

1 10 100 1000

10

20

30

40

50ODR curves of XN-CaO compounds at 165 oC

XN-Ca1.0

XN-Ca1.5

XN-Ca0.5

XN-Ca3.0

XN-Ca2.0

XN-Ca5.0

XN-Ca4.0

Tor

que

(dN

.m)

Time (min)

XNBR

Figure 4.11 ODR curves of XNBR cured with calcium oxide at 165 oC. (log-log scale)

Page 87: Akron 1158067153

69

0 50 100 150 200 250

20

40

60

80

100XN-Ch5.0XN-Ch4.0

XN-Ch3.0

XN-Ch2.0

XN-Ch1.5

XN-Ch0.5

XN-Ch1.0

Tor

que

(dN

.m)

Time (min)

XN-Ca(OH)2 compounds : 165 oC

XNBR

Figure 4.12 ODR curves of XNBR cured with calcium hydroxide at 165 oC. (linear scale)

Page 88: Akron 1158067153

70

1 10 1006789

10

20

30

40

50

60708090

100T

orqu

e (d

N.m

)

Time (min)

XN-Ca(OH)2 compounds : 165 oC

XN-Ch3.0XN-Ch4.0

XN-Ch5.0

XN-Ch2.0

XN-Ch1.5

XN-Ch1.0

XNBR

XN-Ch0.5

Figure 4.13 ODR curves of XNBR cured with calcium hydroxide at 165 oC. (log-log scale)

Page 89: Akron 1158067153

71

There is an induction period, which is dependent on the concentration of Ca(OH)2.

Scorch time reduces, while cure rate increases with increased amounts of Ca(OH)2. A

sharp increase in torque indicates that cure reactions are fast. Maximum torque increases

essentially with increasing Ca(OH)2 concentrations up 1.5x stoichiometry, with little

change thereafter, with excess amounts (2.0x – 5.0x stoichiometric amount).

4.1.5 XNBR-BaO Compositions

ODR curves of XNBR-BaO compounds at 165 oC are shown in Figures 4.14

(linear) and 4.15 (log-log), respectively. At short times, torque of the raw XNBR is

higher than that of compounds containing BaO up to 1.5x stoichiometry. Torque of XN-

Ba0.5 remains lower than that of the neat XNBR for the entire heating time. Similar to

other systems, thermal crosslinking is inhibited by the presence of curatives. The increase

in torque for XN-Ba0.5 with heating time suggests the formation of salts, but these salts

do not yield strong crosslinks, resulting in lower torque than the neat XNBR. It is

important to note that BaO particles are not easily dissolved in the XNBR matrix, and

still remain visible as large particles in all green compounds. Cure rate and stiffness

increase with increasing BaO concentration.

4.2 Crosslink Density Measurements

4.2.1 Thermally-Cured XNBR

Volume fraction (Vr) of the rubber in the swollen gel, sol content, and crosslink

density of thermally cured XNBR at different cure times are shown in Table 4.1, and the

plot of Vr and sol content as a function of cure time is given in Figure 4.16.

Page 90: Akron 1158067153

72

0 50 100 150 200 250

10

20

30

40

50

60

70

80

90Testing temperature : 165 oC

XNBR

XN-Ba5.0

XN-Ba4.0

XN-Ba3.0

XN-Ba2.0

XN-Ba1.5

XN-Ba1.0

Tor

que

(dN

.m)

Time (min)

XN-Ba0.5

Figure 4.14 ODR curves of XNBR cured with barium oxide at 165 oC. (linear scale)

Page 91: Akron 1158067153

73

1 10 100789

10

20

30

40

50

60

708090

100Testing temperature : 165 oC

Tor

que

(dN

.m)

Time (min)

XN-Ba5.0

XN-Ba4.0

XN-Ba3.0

XN-Ba2.0

XN-Ba1.5

XN-Ba1.0

XNBR

XN-Ba0.5

Figure 4.15 ODR curves of XNBR cured with barium oxide at 165 oC. (log-log scale)

Page 92: Akron 1158067153

74

Table 4.1 Volume fraction of rubber (Vr), sol content, and crosslink density of the raw XNBR cured at 165 oC

Cure time (min) Property

60 120 240 500 1000

Vr 0.0147

± 0.0017 0.0389

± 0.0007 0.0711

± 0.0007 0.0827

± 0.0002 0.0962

± 0.0025

Sol content (%) 71.1 ± 1.7

42.6 ± 0.1

22.8 ± 0.2

16.3 ± 0.1

11.8 ± 0.6

ν (x 105 mol/cm3)*

0.47 ± 0.05

0.69 ± 0.07

3.63 ± 0.04

4.45 ± 0.15

6.38 ± 0.10

* Determined by near equilibrium stress-strain measurement

Vr and crosslink density increase with the increased cure times, while sol content

decreases, suggesting self-crosslinking of the raw XNBR, possibly by anhydride

formation. Brown6, 8 suggested that carboxylic rubbers are capable of self-crosslinking to

form anhydride linkages. However, rather severe conditions were required. Lee and

coworkers105 reported that copolymers of ethylene and methacrylic acid containing

various acid contents can form anhydride structures when heated above 140 oC. However,

Vr increases markedly with increasing cure times from 60 to 240 min, and slightly

thereafter. This may be due to existence of equilibrium (eq. 25).106

R C

O

C

O

OH2 C

O

OR R H2O (25)

R = polymer chain

+

Page 93: Akron 1158067153

75

0 200 400 600 800 1000 12000.00

0.02

0.04

0.06

0.08

0.10

0.12

10

20

30

40

50

60

70

80Raw XNBR thermally cured at 165 oC

Vr

Cure time (min)

Vr

Sol content

Sol

con

tent

(%)

Figure 4.16 Vr and sol content of thermally cured XNBR as a function of cure time.

Page 94: Akron 1158067153

76

4.2.2 XNBR-MgO Vulcanizates

Results from equilibrium swelling and stress-strain measurements of XNBR cured

with different magnesium oxides are given in Table 4.2. Figure 4.17 shows the plot of Vr

and sol content against MgO concentration. For vulcanizates containing MgO, Vr

increases with increased concentration of MgO. However, vulcanizates with 0.5x

stoichiometric amount of MgO have lower Vr than the neat XNBR. The swollen gels of

these samples were soft and fragile, and did not maintain their original shape. They lay

flat on the surface of aluminum pans, suggesting weak crosslinking. This is consistent

with ODR results in that the torque of the neat XNBR is higher than those of XN-

MgA0.5, XN-MgB0.5, and XN-MgC0.5. In these vulcanizates, a basic magnesium

hydroxycarboxylate salt is largely formed, and is not expected to give efficient crosslinks.

This salt is a product of the first neutralization step (equation 22), which is an intrachain

reaction, and is expected to be very mobile at curing temperature. Swelling levels of these

samples are much higher than that of the thermally cured XNBR sample. Sol contents of

vulcanizates containing 0.5x stoichiometric amount of MgO are much higher than that of

the neat XNBR, indicating lesser amounts of rubber molecules bound to the networks. Vr

increases markedly with the increased amounts of MgO up to 2.0x stoichiometry, and

changes little thereafter. Apparently, strong rubber networks are obtained when the MgO

concentration is at least 2.0x stoichiometric amounts. Sol content decreases enormously

with the increased concentration of MgO up to 1.5x stoichiometry, and changes little

thereafter, suggesting that at this level of MgO large amounts of rubber are bound to the

networks. High surface area MgO give a slightly higher Vr than do low surface area ones,

but the values are comparable.

Page 95: Akron 1158067153

77

Table 4.2 Volume fraction (Vr) of rubber, sol content, and crosslink density (ν) of XNBR cured with different magnesium oxides (120 min at 165 oC)

Property XNBR XN-MgA0.5

XN-MgA1.0

XN-MgA1.5

XN-MgA2.0

XN-MgA3.0

XN-MgA4.0

XN-MgA5.0

Vr 0.0389

± 0.0007

0.0275±

0.0005

0.0512 ±

0.0003

0.0758 ±

0.0019

0.1028 ±

0.0002

0.1130 ±

0.0010

0.1136 ±

0.0014

0.1223 ±

0.0008

Sol content (%)

42.6 ± 0.1

64.5 ± 0.3

39.6 ± 2.0

8.8 ± 0.3

6.4 ± 0.1

5.6 ± 0.1

5.2 ± 0.2

5.2 ± 0.1

ν (x 105 mol/cm3)*

0.69 ± 0.07

1.46 ± 0.36

4.31 ± 0.21

6.91 ± 0.30

11.5 ± 0.5

15.9 ± 0.6

18.6 ± 0.5

20.0 ± 0.8

Property XNBR XN-MgB0.5

XN-MgB1.0

XN-MgB1.5

XN-MgB2.0

XN-MgB3.0

XN-MgB4.0

XN-MgB5.0

Vr 0.0389

± 0.0007

0.0231 ±

0.0014

0.0389 ±

0.0005

0.0849 ±

0.0040

0.1098 ±

0.0007

0.1105 ±

0.0053

0.1104 ±

0.0027

0.1134 ±

0.0015

Sol content (%)

42.6 ± 0.1

73.0 ± 1.7

52.9 ± 0.5

7.2 ± 0.7

6.4 ± 0.1

5.6 ± 0.2

4.8 ± 0.1

4.6 ± 0.1

ν (x 105 mol/cm3)*

0.69 ± 0.07

0.57 ± 0.06

1.50 ± 0.17

2.81 ± 0.59

7.39 ± 0.24

9.00 ± 0.86

15.7 ± 0.7

18.1 ± 0.5

Property XNBR XN-MgC0.5

XN-MgC1.0

XN-MgC1.5

XN-MgC2.0

XN-MgC3.0

XN-MgC4.0

XN-MgC5.0

Vr 0.0389

± 0.0007

0.0193 ±

0.0005

0.0319 ±

0.0007

0.0828 ±

0.0019

0.1011 ±

0.0047

0.0920 ±

0.0082

0.0900 ±

0.0018

0.1081 ±

0.0007

Sol content (%)

42.6 ± 0.1

72.8 ± 0.2

63.3 ± 1.0

9.5 ± 0.3

7.7 ± 0.3

7.1 ± 0.1

6.4 ± 0.1

6.3 ± 0.3

ν (x 105 mol/cm3)*

0.69 ± 0.07

0.23 ± 0.02

0.97 ± 0.19

2.65 ± 0.40

4.44 ± 0.33

8.57 ± 0.32

15.1 ± 0.34

19.5 ± 0.25

* Determined by near equilibrium stress-strain measurement

Page 96: Akron 1158067153

78

0 1 2 3 4 5 6

0.02

0.04

0.06

0.08

0.10

0.12

0.14

0

10

20

30

40

50

60

70

80

90

XN-MgA XN-MgB XN-MgC

opened is Vr

closed is sol content

Vr

Amount of MgO/ Stoichiometric Amount

XNBR-MgO compounds cured 120 min at 165 oC

Vr

Sol content

Sol

con

tent

(%)

Figure 4.17 Vr and sol content of MgO-cured XNBR as a function of MgO concentration.

Page 97: Akron 1158067153

79

4.2.3 XNBR-Peroxide Vulcanizates

Vr, sol content, and crosslink density of XNBR-peroxide vulcanizates are given in

Table 4.3. The plot of Vr and sol content as a function of peroxide content is shown in

Figure 4.18. As expected, Vr and crosslink density increase, while the sol content

decreases with increased peroxide content. The shape of the plot between Vr and peroxide

amount is very similar to Figure 4.9, indicating a decline of crosslink efficiency.

Table 4.3 Volume fraction (Vr) of rubber, sol content, and crosslink density of XNBR cured with dicumyl peroxide 60 min at 165 oC

Property XNBR XN-P0.25

XN-P0.5

XN-P0.75

XN-P1.0

XN-P1.5

XN-P2.0

XN-P3.0

Vr 0.0147

± 0.0017

0.0410±

0.0010

0.0713 ±

0.0015

0.0905 ±

0.0007

0.1098 ±

0.0008

0.1370 ±

0.0007

0.1552 ±

0.0005

0.1869 ±

0.0003

Sol content (%)

71.1 ± 1.7

34.7 ± 0.1

19.6 ± 0.3

13.9 ± 0.1

10.4 ± 0.1

7.7 ± 0.1

6.0 ± 0.1

5.2 ± 0.2

ν(x 105 mol/cm3)*

0.47 ± 0.05

1.83 ± 0.17

3.38 ± 0.24

4.81 ± 0.40

5.85 ± 0.47

12.4 ± 0.5

13.1 ± 0.4

23.4 ± 0.3

* Determined by near equilibrium stress-strain measurement

4.2.4 XNBR-CaO Vulcanizates

Results from equilibrium swelling and stress-strain measurements are given in

Table 4.4, and the plot of Vr and sol content against CaO content is shown in Figure 4.19.

Vr and crosslink density of the neat XNBR is slightly higher than for XN-Ca vulcanizates,

while the sol content is slightly less. However, Vr and the sol content of all compounds

are comparable. Apparently, curing in all the compounds is very similar; that is thermal

crosslinking.

Page 98: Akron 1158067153

80

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.50.00

0.04

0.08

0.12

0.16

0.20

0

10

20

30

40

50

60

70

80

Vr

Amount of peroxide (phr)

Vr

XNBR-Peroxide compounds cured 60 min at 165 oC

Sol content

Sol

con

tent

(%)

Figure 4.18 Vr and sol content of XNBR-Peroxide vulcanizates as a function of peroxide concentration.

Page 99: Akron 1158067153

81

Table 4.4 Volume fraction (Vr) of rubber, sol content, and crosslink density of XNBR cured with calcium oxide 1000 min at 165 oC

Property XNBR XN-Ca0.5

XN-Ca1.0

XN-Ca1.5

XN-Ca2.0

XN-Ca3.0

XN-Ca4.0

XN-Ca5.0

Vr 0.0962

± 0.0025

0.0883±

0.0011

0.0806 ±

0.0011

0.0822 ±

0.0062

0.0849 ±

0.0021

0.0833 ±

0.0007

0.0869 ±

0.0004

0.0803 ±

0.0004

Sol content (%)

11.8 ± 0.6

12.9 ± 0.1

14.6 ± 0.3

15.1 ± 2.4

15.1 ± 1.3

13.5 ± 0.5

13.8 ± 0.1

14.9 ± 0.4

ν (x 105 mol/cm3)*

6.38 ± 0.10

6.33 ± 0.17

5.13 ± 0.08

4.20 ± 0.12

5.94 ± 0.17

4.37 ± 0.26

4.47 ± 0.36

4.21 ± 0.14

* Determined by near equilibrium stress-strain measurement

A lower Vr in XN-Ca vulcanizates is probably because self-crosslinking is

prohibited by the presence of CaO. At first CaO is expected to have a similar impact on

properties of the XNBR as ZnO. However, CaO curing does not lead to ionic crosslinks,

which give high tensile properties. Starmer25 classified CaO as an inactive material,

having a conflicting effect on the properties of the XNBR when compared to ZnO, an

active one.

4.2.5 XNBR-Ca(OH)2 Vulcanizates

Vr, sol content, and crosslink density of XNBR-Ca(OH)2 vulcanizates are given in

Table 4.5. The plot of Vr and sol content versus Ca(OH)2 content is shown in Figure 4.20.

Page 100: Akron 1158067153

82

0 1 2 3 4 5 60.00

0.04

0.08

0.12

0.16

0.20

0

5

10

15

20

25

Vr

Amount of CaO/Stoichiometric Amount

XN-Ca compounds cured 1000 min at 165 oC

Sol content

Vr

Sol

con

tent

(%)

Figure 4.19 Vr and sol content of XNBR-CaO vulcanizates as a function of CaO concentration.

Page 101: Akron 1158067153

83

Table 4.5 Volume fraction (Vr) of rubber, sol content, and crosslink density of XNBR cured with calcium hydroxide 240 min at 165 oC

Property XNBR XN-Ch0.5

XN-Ch1.0

XN-Ch1.5

XN-Ch2.0

XN-Ch3.0

XN-Ch4.0

XN-Ch5.0

Vr 0.0711

± 0.0007

0.0193±

0.0039

0.0316 ±

0.0011

0.0588 ±

0.0012

0.0717 ±

0.0008

0.0777 ±

0.0015

0.0880 ±

0.0004

0.0963 ±

0.0004

Sol content (%)

22.8 ± 0.2

66.4 ± 1.4

41.7 ± 0.3

12.8 ± 0.9

9.4 ± 0.2

7.2 ± 0.6

6.3 ± 0.2

5.4 ± 0.4

ν (x 105 mol/cm3)*

3.63 ± 0.04

5.34 ± 0.08

14.7 ± 0.9

29.1 ± 0.8

30.6 ± 1.2

32.1 ± 0.8

32.5 ± 0.1

26.7 ± 2.1

* Determined by near equilibrium stress-strain measurement

Vr and crosslink density increase with increased Ca(OH)2 content, while the sol

content decreases. Obviously, Vr of the neat XNBR is higher than those of vulcanizates

containing Ca(OH)2 up to 1.5x stoichiometry, but similar to that of the XN-Ch2.0

vulcanizate. The swollen gels of XN-Ch0.5 and XN-Ch1.0 were soft and fragile, and did

not maintain their shape, and lay flat on the aluminum pan surface, suggesting weak

crosslinking. Vulcanizates containing Ca(OH)2 3.0x to 5.0x stoichiometry have higher Vr

than the neat XNBR, indicating strong crosslinking. The sol content decreases sharply

with increasing amount of Ca(OH)2 up to 1.5x stoichiometry, and then changes slightly

with a great excess of Ca(OH)2, suggesting that most of the rubber molecules are bound

to the rubber network. Apparently, strong crosslinking results when the amount of

Ca(OH)2 is at least twice stoichiometry.

Page 102: Akron 1158067153

84

0 1 2 3 4 5 60.00

0.02

0.04

0.06

0.08

0.10

0.12

0

10

20

30

40

50

60

70

80

Vr

Amount of Ca(OH)2/Stoichiometric Amount

XN-Ca(OH)2 compounds cured 240 min at 165 oC

Vr

Sol content

Sol

con

tent

(%)

Figure 4.20 Vr and sol content of XNBR-Ca(OH)2 vulcanizates as a function of Ca(OH)2 concentration.

Page 103: Akron 1158067153

85

4.2.6 XNBR-BaO Vulcanizates

Results from equilibrium swelling and stress-strain testes are given in Table 4.6.

Figure 4.21 shows the influence of the amounts of BaO on Vr and the sol content of

XNBR-BaO vulcanizates. As in the cases of MgO and Ca(OH)2, Vr and crosslink density

increase with increasing BaO content, with a corresponding decrease in sol content. Vr

increases greatly with the increased BaO content up 2.0x stoichiometry. The sol content

decreases with increasing BaO amounts up to 2.0x stoichiometry. It is interesting to note

that for all the XNBR-BaO vulcanizates, Vr is less than that of the neat XNBR. However,

crosslink density of XNBR containing BaO at least 1.5x stoichiometry, determined by

near equilibrium stress-strain measurement, is higher than that of the neat XNBR. This

suggests that solvent resistance of ionic crosslinks formed is worse than covalent

crosslinks.

Table 4.6 Volume fraction (Vr) of rubber, sol content, and crosslink density of XNBR cured with barium oxide 240 min at 165 oC

Property XNBR XN-Ba0.5

XN-Ba1.0

XN-Ba1.5

XN-Ba2.0

XN-Ba3.0

XN-Ba4.0

XN-Ba5.0

Vr 0.0711

± 0.0007

0.0344±

0.0015

0.0441 ±

0.0005

0.0491 ±

0.0005

0.0552 ±

0.0087

0.0588 ±

0.0002

0.0649 ±

0.0017

0.0684 ±

0.0010

Sol content (%)

22.8 ± 0.2

38.1 ± 2.5

23.1 ± 0.2

17.4 ± 0.1

13.7 ± 0.1

9.9 ± 0.1

6.2 ± 0.2

4.2 ± 0.8

ν (x 105 mol/cm3)*

3.63 ± 0.04

2.12 ± 0.21

3.09 ± 0.47

6.17 ± 0.04

8.59 ± 0.31

9.29 ± 0.65

11.0 ± 0.7

11.5 ± 0.8

* Determined by near equilibrium stress-strain measurement

Page 104: Akron 1158067153

86

0 1 2 3 4 5 60.02

0.03

0.04

0.05

0.06

0.07

0.08

0.09

0

10

20

30

40

50XNBR-BaO compounds cured 240 min at 165 oC

Vr

Amount of BaO/Stoichiometric Amount S

ol c

onte

nt (%

)

Figure 4.21 Vr and sol content of XNBR-BaO vulcanizates as a function of BaO concentration.

Page 105: Akron 1158067153

87

4.2.7 Comparison among Metal Compounds

Vr and sol content of XNBR vulcanized by various metal compounds are shown

as a function of metal compound concentration in Figures 4.22 and 4.23, respectively. Vr

and sol content of XNBR-CaO vulcanizates (cured for 1000 min at 165 oC) show little

change with increasing CaO contents, and are comparable to those of the neat XNBR. In

these vulcanizates, curing is very similar, and occurs by thermal coupling of carboxyl

groups to from covalent anhydride bridges. Presence of CaO particles may somehow

inhibit thermal crosslinking. Therefore, Vr of XNBR-CaO samples is slightly lower than

that of the neat XNBR.

XN-Ba0.5 has a lower sol content and slightly higher Vr than those of XN-Ch0.5

and XN-MgA0.5. This may be due to the effect of particle size. As mentioned before, all

XN-Ba vulcanizate sheets contained visible particles. These may lead to larger amounts

of carboxyl groups that have not reacted with BaO. These carboxyl groups can undergo

self-coupling to form anhydride crosslinks. Another possibility is that ionic salts formed

in these samples do not give efficient crosslinking, therefore anhydride links will be more

important. For samples containing 1.0x stoichiometry, MgO gives higher Vr than BaO

and Ca(OH)2. We will show later that in XN-MgA1.0 all carboxyl groups are essentially

neutralized, and both magnesium hydroxycarboxylate (inefficient crosslink) and

magnesium carboxylate (efficient crosslink) are formed, while in XN-Ch1.0 and XN-

Ba1.0 there are certain amounts of carboxyl groups, which have not reacted with metal

compounds.

Page 106: Akron 1158067153

88

0 1 2 3 4 5 6

0.02

0.04

0.06

0.08

0.10

0.12

0.14

Vr

Amount of metal compounds/Stoichiometric amount

XN-MgO

XN-Ca(OH)2

XN-CaO

XN-BaO

Figure 4.22 Vr of XNBR vulcanized by various metal compounds as a function of concentration.

Page 107: Akron 1158067153

89

0 1 2 3 4 5 60

10

20

30

40

50

60

70

80

Amount of metal compounds/Stoichiometric amount

XN-CaO XN-Ca(OH)2

XN-BaO XN-MgO

Sol c

onte

nt (%

)

Figure 4.23 Sol content of XNBR vulcanized by various metal compounds as a function of concentration.

Page 108: Akron 1158067153

90

With increasing amounts of metal compounds up to 1.5x stoichiometry, sol

contents decrease, but change little thereafter. It is interesting to note that sol contents are

similar among XN-Ba, XN-Ch, and XN-MgA systems above this concentration. This

suggests that the amount of rubber molecules bound into a network is very similar in

these systems. Vr markedly increase with increased concentration of metal compounds up

to 2.0x stoichiometry, and slightly increases with further great excess. Vr and sol content

obviously indicate that the amount of rubber bound into a network is similar in these

three systems, but they differ in the degree of swelling. This suggests a difference in the

strength of the crosslinks in each system, and a possible relation between the strength of

crosslinks and the size of metal ions. The strength of ionic crosslinks is in the following

order: XNBR-MgO > XNBR-Ca(OH)2 > XNBR-BaO. Table 4.7 shows the effective

ionic radii of Mg++, Ca++, and Ba++ with various coordination numbers.107 At the same

coordination number, ionic radii are ranked as followed: Ba++ > Ca++ > Mg++.

Table 4.7 The effective ionic radii of Mg++, Ca++, and Ba++ ions with various coordination numbers107

Type of Ion Coordination Number

Ionic Radii (Å)

Mg++ 4 6 8

0.71 0.86 1.03

Ca++ 6 8

1.14 1.26

Ba++ 6 8

1.49 1.56

Page 109: Akron 1158067153

91

The Coulombic force, F, between an anion of charge, q-, and a cation of charge,

q+, is given by

20 r4qqFεεπ

=−+

(26)

where r is the distance between the centers of the ions, ε is the dielectric constant of the

medium, and ε0 is the permittivity of free space (8.8542 x 10-12 C2/N.m2). The attractive

force is directly proportional to the product of ionic charges, and varies inversely with the

square of the distance between them. Clearly, swelling behaviors of XNBR-MgO,

XNBR-Ca(OH)2, and XNBR-BaO systems can be well explained by the classical

Coulomb law. Tant and coworkers84 made similar observations in IA and IIA metal-

neutralized carboxylate telechelic polyisoprene. They also suggested that this simple rule

holds only within the particular group of the periodic table, but cannot be applied across

groups. Bagrodia and Wilkes108 commented that not only does the nature of the cation

play a role in determining ionomer properties, but also the electronic configuration of the

cation, which governs its covalent characteristics.

4.3 Tensile Properties

4.3.1 Thermally Cured XNBR

Stress-strain curves of thermally cured XNBR are shown in Figure 4.24, and

tensile properties, 300% modulus, tensile strength, and breaking strain, are plotted as a

function of cure time in Figure 4.25.

Page 110: Akron 1158067153

92

0 200 400 600 800 1000 1200 1400 16000

2

4

6

8

10

120 min

240 min

500 min

x

x

x

Stre

ss (M

Pa)

Strain (%)

Thermally cured raw XNBR at 165 oC

x

1000 min

Figure 4.24 Stress-strain curves of thermally cured XNBR.

Page 111: Akron 1158067153

93

0 200 400 600 800 10000

1

6

7

8

9

10

11

600

800

1000

1200

1400

1600

300%

Mod

ulus

or

Ten

sile

stre

ngth

(MPa

)

Cure time (min)

Raw XNBR thermally cured at 165 oC

Tensile strength

Elongation at break

300% Modulus

Elo

ngat

ion

at b

reak

(%)

Figure 4.25 Tensile properties of thermally cured XNBR as a function of cure time.

Page 112: Akron 1158067153

94

Modulus and tensile strength increase rapidly, while the ultimate strain decreases

with increased cure time from 60 to 240 min. This is due to the formation of anhydride

crosslinks upon heating. Further increase in cure times (500 and 1000 min) results in little

improvement in tensile properties. This may be due to reaching an equilibrium in

anhydride formation (equation 25),106 which will limit the amount of anhydride crosslinks

formed.

4.3.2 XNBR-MgO Vulcanizates

Stress-strain curves of XN-MgA, XN-MgB, and XN-MgC vulcanizates (cured

120 min at 165 oC) are given in Figures 4.26, 4.27, and 4.28, respectively. Similar

behaviors are observed in all three systems. Tensile modulus and strength are very much

improved, while the breaking strain decreases with the increased MgO content up to 1.5x

to 2.0x stoichiometry, with little change for a great excess of MgO. Tensile results are

consistent with those from ODR and swelling measurements.

Figure 4.29 shows tensile properties of XN-MgA, XN-MgB, and XN-MgC

vulcanizates as a function of MgO contents. 300% Modulus and tensile strength increase

with increasing MgO content up to 1.5x to 2.0x stoichiometry, and slightly change

thereafter. Brown5, 6 obtained similar results on ZnO-cured XNBR, and so did Sato on

ZnO-cured XSBR.64 Breaking strain markedly decreases with increasing MgO content up

to 1.0x to 1.5x stoichiometry, and changes little after that. A tensile strength of about 48

to 52 MPa with an ultimate strain of 500 to 600% is exceptional high, suggesting that

MgO-cured XNBR is a self-reinforced system. However, optimum properties require an

MgO content of at least twice stoichiometry.

Page 113: Akron 1158067153

95

0 400 800 1200 16000

10

20

30

40

50

60

xxx

x

x

x

x

XN-MgA5.0

XN-MgA4.0

XN-MgA3.0

XN-MgA2.0

XN-MgA1.5

XN-MgA1.0

XN-MgA0.5

Stre

ss (M

Pa)

Strain (%)

XNBR

XN-MgA vulcanizates cured 120 min at 165 oC

x

Figure 4.26 Stress-strain curves of XN-MgA vulcanizates (cured 120 min at 165 oC).

Page 114: Akron 1158067153

96

0 400 800 1200 16000

10

20

30

40

50

60

xxx

x

x

x

x

XN-MgB5.0

XN-MgB4.0

XN-MgB3.0

XN-MgB2.0

XN-MgB1.5

XN-MgB1.0

XN-MgB0.5

XNBR

Stre

ss (M

Pa)

Strain (%)

XN-MgB vulcanizates cured 120 min at 165 oC

x

Figure 4.27 Stress-strain curves of XN-MgB vulcanizates (cured 120 min at 165 oC).

Page 115: Akron 1158067153

97

0 400 800 1200 16000

10

20

30

40

50

60

XN-MgC4.0

XN-MgC5.0

XN-MgC3.0

XN-MgC2.0

XN-MgC1.5

XN-MgC1.0

XNBRXN-MgC0.5

xx

x

x

x

x

x

Stre

ss (M

Pa)

Strain (%)

x

XN-MgC vulcanizates cured 120 min at 165 oC

Figure 4.28 Stress-strain curves of XN-MgC vulcanizates (cured 120 min at 165 oC).

Page 116: Akron 1158067153

98

0 1 2 3 4 5 6

1

10

100

400

600

800

1000

1200

1400

1600

Elongation at break

XN-MgB

300%

Mod

ulus

or

Ten

sile

stre

ngth

(MPa

)

Amount of MgO/Stiochiometric Amount

XNBR-MgO vulcanizates cured 120 min at 165 oCXN-MgA Tensile strength

XN-MgC

XN-MgA

XN-MgB XN-MgC

300% Modulus

XN-MgA XN-MgB XN-MgC

Elo

ngat

ion

at b

reak

(%)

Figure 4.29 Tensile properties of XNBR cured with different magnesium oxides (cured 120 min at 165 oC).

Page 117: Akron 1158067153

99

It may be deduced from the tensile results that carboxyl groups are completely

neutralized at about 2.0x stoichiometry. High tensile strength and modulus is due to

aggregation of ionic crosslinks to form a biphasic reinforced structure as suggested by

Tobolsky and coworkers.53 Ionic aggregates may function both as multifunctional

crosslinks and as reinforcing filler particles. Additionally, Cooper72-74 suggested that high

tensile strength may be due to an interchange between ionic crosslinks under mechanical

stress. This mechanism will prevent the development of local stress concentration which

can lead to catastrophic failure. The required amount (at least 2.0x stoichiometry) of

MgO to gain optimum properties proves that the classical neutralization (eq. 7) for

divalent metals, in which one mole of metal oxide neutralizes two moles of carboxyl

groups, is incorrect and we will show later that it also cannot hold for other metal

compounds studied. Many researchers have suggested that a basic salt, – COOMOH, may

form in carboxylic rubbers neutralized by divalent metal oxides, and that its polar nature

can lead to strong intermolecular interactions.5-8, 59, 72-74, 81, 82

Apparently, the amount of MgO is a major factor governing the tensile properties

of vulcanizates up to the point (less than 2.0x stoichiometry) that all carboxyl groups are

completely neutralized and a strong neutral salt, – COOMOOC –, is formed. After that,

the effect of concentration is not significant. Specific surface area is not an important

factor in determining the tensile properties of vulcanizates (Figure 4.29), although high

specific surface area gives slightly better tensile properties. It appears that specific

surface area has a great impact only on the cure behavior.

Page 118: Akron 1158067153

100

4.3.3 XNBR-Peroxide Vulcanizates

Stress-strain curves of peroxide-cured XNBR are given in Figure 4.30, and the

plot of tensile properties against peroxide content is shown in Figure 4.31. The neat

XNBR heated for 60 min at 165 oC has an elongation greater than 1600 % with tensile

stress greater than 4.2 MPa. A maximum tensile strength of 8.33 MPa with breaking

strain of about 1400 % is obtained in a vulcanizate containing small amount of dicumyl

peroxide (0.25 phr). With increasing curative amounts the tensile strength drops. The

ultimate strain decreases, while modulus increases linearly with the amount of peroxide.

In the rupture of rubber vulcanizates, a portion of the input energy is stored elastically

and released upon crack propagation. The rest of the energy is lost in internal dissipative

processes, such as chain motion. At high crosslink levels, chain motions are restricted,

and not much energy is dissipated. This results in brittle fracture at low strain.109

4.3.4 XNBR-CaO Vulcanizates

Figure 4.32 shows the stress-strain curves of XN-Ca vulcanizates (cured 1000 min

at 165 oC). They are very similar to that of the neat XNBR. The effect of CaO content on

the tensile properties of XN-Ca vulcanizates is given in Figure 4.33. Tensile strengths (7

to 8 MPa) of XN-Ca samples are slightly less than that of the neat XNBR (about 9 MPa),

but not by much. The 300% modulus and elongation at break of XN-Ca vulcanizates are

approximately the same as those of the raw, thermally cured XNBR. This is indirect

evidence that curing of the raw XNBR and CaO-containing XNBR are similar, and that

salt formation does not occur.

Page 119: Akron 1158067153

101

0 200 400 600 800 1000 1200 1400 1600 18000

2

4

6

8

10

x x

x

x

x

x

XNBRnot breakSt

ress

(MPa

)

Strain (%)

XN-Peroxide vulcanizates cured 60 min at 165 oC

XN-P0.50XN-P0.25

XN-P0.75

XN-P1.0

XN-P1.5

XN-P2.0XN-P3.0

x

Figure 4.30 Stress-strain curves of XN-peroxide vulcanizates (cured 60 min at 165 oC).

Page 120: Akron 1158067153

102

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.50

1

2

3

4

5

6

7

8

9

10

0

200

400

600

800

1000

1200

1400

1600

300%

Mod

ulus

or

Ten

sile

stre

ngth

(MPa

)

Amount of dicumyl peroxide (phr)

XN-Peroxide compounds cured 60 min at 165 oC

Tensile strength

Elongation at break

300%Modulus

Elo

ngat

ion

at b

reak

(%)

Figure 4.31 Tensile properties of XN-peroxide vulcanizates (cured 60 min at 165 oC).

Page 121: Akron 1158067153

103

These XNBR-CaO systems do not exhibit self-reinforcement. It appears that their

tensile properties are largely determined by covalent crosslinks rather than by ionic

crosslinks. In fact, Starmer25 classified CaO as an inactive material for XNBR. It is not

yet understood why CaO does not provide ionic crosslinks in XNBR.

4.3.5 XNBR-Ca(OH)2 Vulcanizates

Stress-strain curves of XNBR cured with Ca(OH)2 are given in Figure 4.34, and

the dependence of tensile properties on the concentration of curatives is shown in Figure

4.35. Tensile strength and 300% modulus increase markedly with increased amounts of

Ca(OH)2 up to 1.5x stoichiometry, with little change thereafter. Elongation at rupture

decreases with increasing Ca(OH)2 contents up to 1.0x to 1.5x stoichiometry, and then

saturates at about 500%. The maximum tensile strength of about 50 MPa is much greater

than that of the raw XNBR (about 8 MPa), indicating that Ca(OH)2-cured XNBR exhibit

self-reinforcement like MgO-vulcanized XNBRs. Clearly, ionic crosslinks are formed in

these vulcanizates, and aggregate to form hard domains which act as multifunctional

crosslinks and reinforcing structures. This accounted for the high tensile strength and

modulus of XNBR cured with Ca(OH)2.53, 56, 71 The ability for crosslinks to interchange

under mechanical stress will also prevent local stress concentration. These mechanisms

are reasons for high tensile modulus and strength of XNBR-Ca(OH)2 vulcanizates.

Page 122: Akron 1158067153

104

0 100 200 300 400 500 600 700 8000

2

4

6

8

10

XNCa3.0

XNCa4.0

XNCa5.0

XNCa1.5

XNCa2.0

XNCa1.0

XN-Ca0.5

xxx

xx

x

x

Stre

ss (M

Pa)

Strain (%)

x

XNBR

XN-Ca vulcanizates cured 1000 min at 165 oC

Figure 4.32 Stress-strain curves of XN-Ca vulcanizates (cured 1000 min at 165 oC).

Page 123: Akron 1158067153

105

0 1 2 3 4 5 60

1

2

3

4

5

6

7

8

9

10

500

600

700

800

900

1000

300%

Mod

ulus

or

Ten

sile

stre

ngth

(MPa

)

Amount of CaO/Stoichiometric Amount

XN-CaO compounds cured 1000 min at 165 oC

Tensile strength

Elongation at break

300% Modulus

Elo

ngat

ion

at b

reak

(%)

Figure 4.33 Tensile properties of XN-Ca vulcanizates (cured 1000 min at 165 oC).

Page 124: Akron 1158067153

106

0 200 400 600 800 10000

10

20

30

40

50

60

XN-Ch5.0

XN-Ch4.0

XN-Ch3.0

XN-Ch2.0XN-Ch1.5

XN-Ch1.0

xxx

x

x

x

xXN-Ch0.5

Stre

ss (M

Pa)

Strain (%)

XNBRx

XN-Ca(OH)2 compounds cured 240 min at 165 oC

Figure 4.34 Stress-strain curves of XN-Ch vulcanizates (cured 240 min at 165 oC).

Page 125: Akron 1158067153

107

0 1 2 3 4 5 6

1

10

100

400

500

600

700

800

900

1000

300%

Mod

ulus

or

Ten

sile

stre

ngth

(MPa

)

Amount of Ca(OH)2/Stoichiometric Amount

XN-Ca(OH)2 compounds cured 240 min at 165 oC

Tensile strength

300% Modulus

Elongation at break

Elo

ngat

ion

at b

reak

(%)

Figure 4.35Tensile properties of XN-Ch vulcanizates (cured 240 min at 165 oC).

Page 126: Akron 1158067153

108

4.3.6 XNBR-BaO Vulcanizates

Figure 4.36 shows stress-strain curves of XNBR-BaO vulcanizates cured for 240

min at 165 oC. Similar to MgO- and Ca(OH)2-cured samples, the tensile moduli increase,

while the breaking strain falls with increased amount of BaO. Dependence of tensile

properties on BaO concentrations is given in Figure 4.37. The 300% Modulus

monotonically increases, while the tensile strength greatly increases with increased BaO

concentration up to twice stoichiometry. The breaking strain decreases with the BaO

content up to 1.0x - 1.5x stoichiometric amounts, and slightly decreases thereafter. The

maximum tensile strength of about 27 MPa at about 500 % breaking strain suggests that

BaO-cured XNBR is also self-reinforcement, like MgO- and Ca(OH)2-cured XNBR.

However, the degree of reinforcement in BaO-cured samples is less than that in MgO-

and Ca(OH)2-vulcanized XNBRs. BaO is not well-dissolved in the XNBR. When the

vulcanized sheets were prepared, BaO particles remained large, visible, and not well-

dispersed. These large particles will act as stress-raisers which magnify applied stresses

and reduce the tensile strength.43

4.3.7 Comparison of Tensile Properties among Metal Compounds

Figure 4.38 shows 300% moduli of XNBR vulcanized with different metal

compounds. The 300% moduli of XNBR-CaO vulcanizates are approximately the same

as that of the raw, thermally cured XNBR. In these systems no ionic crosslinks are

formed, and tensile properties are governed by covalent crosslinks. For the rest of metal

compounds, Ca(OH)2 gives the highest modulus. MgO gives higher modulus than BaO

when the amounts of curatives present are equal or less than 4.0x stoichiometry.

Page 127: Akron 1158067153

109

0 200 400 600 800 10000

5

10

15

20

25

30XN-BaO vulcanizates cured 240 min at 165 oC

XN-Ba5.0

XN-Ba4.0XN-Ba3.0

XN-Ba2.0

XN-Ba1.5

XNBR

XN-Ba1.0

xx

x

x

x

xx

Stre

ss (M

Pa)

Strain (%)

x

XN-Ba0.5

Figure 4.36 Stress-strain curves of XN-Ba vulcanizates (cured 240 min at 165 oC).

Page 128: Akron 1158067153

110

0 1 2 3 4 5 60

5

10

15

20

25

30

35

300

400

500

600

700

800

900

1000

1100

300%

Mod

ulus

or

Ten

sile

stre

ngth

(MPa

)

Amount of BaO/Stoichiometric Amount

XN-BaO vulcanizates cured 240 min at 165 oC

Tensile strength

300% Modulus

Elongation at break

Elo

ngat

ion

at b

reak

(%)

Figure 4.37 Tensile properties of XN-Ba vulcanizates (cured 240 min at 165 oC).

Page 129: Akron 1158067153

111

For XNBR-Ca(OH)2 vulcanizates, strong ionic salts are formed even at a stoichiometric

amounts of Ca(OH)2 (discussed later in the ATR-IR section). Therefore, strong ionic

domains are expected, and the number of effective network chains is high, resulting in

high modulus. A slight increase with excess amounts of Ca(OH)2 may be due to a

hydrodynamic effect. In the case of MgO-cured samples, strong ionic salts are not formed

until the amount of MgO is equal or greater than 2.0x stoichiometry. At low

concentrations (0.5x to 1.0x stoichiometry), the magnesium hydroxycarboxylate salt,

– COOMgOH, is the main product. This type of salt will not give efficient crosslinks,

therefore, the number of effective network chains is low, resulting in a low modulus

when compared to Ca(OH)2-cured vulcanizates. For XNBR cured with BaO, due to

incomplete solubility, neutralization is not complete until 2.0x stoichiometry of BaO is

present; therefore the amount of salts formed is low. The number of effective network

chains is expected to be lower than in MgO- and Ca(OH)2-vulcanized samples, with a

lower modulus.

Tensile strengths of XNBR vulcanized by various metal compounds are compared

in Figures 4.39. The effect of metal compounds on tensile strength is as followed: MgO >

Ca(OH)2 > BaO > CaO. The first three metal compounds result in ionic crosslinks, while

no salt is formed in XNBR cured with CaO. Substantial reinforcement is obtained with

MgO and Ca(OH)2, while BaO gives a moderate effect. As discussed before, high tensile

strength is attributed to aggregation of ionic crosslinks to form reinforcing hard domains

which also function as multifunctional crosslinks.52, 56, 71 Apparently, such reinforcing

domains are obtained when at least 2.0x stoichiometric amounts of metal compounds are

present. Furthermore, interchange between ionic crosslinks will prevent locally high

Page 130: Akron 1158067153

112

stress concentration, and allow all of the whole network chains to bear mechanical load.

However, ionic crosslinks should be strong enough to yield the characteristic high tensile

strength.72-74 From the swelling results, the strength of ionic crosslinks are in the

following order: MgO > Ca(OH)2 > BaO (Figure 4.22). Clearly, the tensile results are in

good agreement with these swelling measurements.

Figure 4.40 shows the ultimate strains of XNBR cured with different metal

compounds. Breaking strains of XNBR-CaO vulcanizates are approximately the same as

for the neat XNBR. As discussed earlier, curing of these systems is similar to that of

thermally cured XNBR. Covalent crosslinks are the main products. The effect of the

other three metal compounds on breaking strains, MgO, Ca(OH)2, and BaO, are very

similar. The ultimate strains decrease with the increased amounts of metal compounds up

to 1.0x to 1.5x stoichiometry, and change little thereafter. In fact, at a stoichiometric

amount of metal compounds, salts are already formed, but apparently, they do not yield

efficient crosslinks, evidenced by low tensile strength. However, association of these

ionic salts can change the topology of rubber networks, causing highly entangled

networks which result in rupture of the rubber at low strains.

It appears that the breaking strains reach a maximum level at lower concentrations

(1.0x to 1.5x stoichiometry) of metal compounds than for tensile strengths (1.5x to 2.0x

stoichiometry). This is indirect evidence that additional network chains are not formed

until 1.5x to 2.0x stoichiometric amounts of metal compounds are present.

Page 131: Akron 1158067153

113

0 1 2 3 4 5 60

5

10

15

20

25

300%

Mod

ulus

(MPa

)

Amount of metal compounds/Stoichiometric amount

XNBR-Ca(OH)2

XNBR-BaO

XNBR-MgO

XNBR-CaO

Cured at 165 oC

Figure 4.38 300% Modulus of the XNBR vulcanized by various metal compounds.

Page 132: Akron 1158067153

114

0 1 2 3 4 5 60

10

20

30

40

50

60

Amount of metal compounds/Stoichiometric amount

Ten

sile

stre

ngth

(MPa

)Cured at 165 oC

XNBR-MgO

XNBR-Ca(OH)2

XNBR-BaO

XNBR-CaO

Figure 4.39 Tensile strength of the XNBR vulcanized by various metal compounds.

Page 133: Akron 1158067153

115

0 1 2 3 4 5 6200

400

600

800

1000

1200

1400

1600

Amount of metal compounds/Stoichiometric amount

Elo

ngat

ion

at b

reak

(%)

Cured at 165 oC

XNBR-CaO

XNBR-MgO

XNBR-BaO

XNBR-Ca(OH)2

Figure 4.40 Elongation at break of the XNBR vulcanized by various metal compounds.

Page 134: Akron 1158067153

116

4.3.8 Comparison between Ionic and Covalent Crosslinks

In this section, the effect of crosslink types on the tensile properties of XNBR is

discussed. Figure 4.41 shows 300% modulus of XNBR vulcanized by various curing

agents, which yield different types of crosslinks. The 300% Moduli of ionically

crosslinked samples (cured by MgO, Ca(OH)2, and BaO) increase with Vr more than for

covalently crosslinked ones (cured by dicumyl peroxide, CaO, and thermal energy). This

is attributed to the aggregation of ionic crosslinks to form hard domains. However, the

effect varies among systems, probably due to a difference in the number of effective

network chains formed. In the case of covalently crosslinked samples, the 300% modulus

increases linearly with Vr, and appears to be independent of the crosslink structures, i.e.,

carbon-carbon crosslinks in XNBR-peroxide systems or anhydride crosslinks in XNBR-

CaO systems and in thermally cured XNBRs.

Tensile strengths of XNBR cured by different agents are given in Figure 4.42 as a

function of Vr. Apparently, the tensile strength of ionically crosslinked rubbers increase

with increased Vr until all the carboxyl groups are completely neutralized, and changes

slightly thereafter. For covalently crosslinked rubbers, the tensile strength decreases with

increased crosslink density. In comparison with covalent crosslinks, ionic crosslinks give

vulcanizates with much higher tensile strength. This is a result of association of ionic

crosslinks to form hard reinforcing domains.53, 56, 71 These ionic crosslinks can

interchange under mechanical stress. This relieving mechanism will prevent locally high

stress concentrations.72-74 In other words, the high tensile strength in ionically crosslinked

rubbers is primarily due to the ability to relax stress. However, ionic crosslinks should be

sufficiently strong to give the characteristic high tensile strength. As mention before,

Page 135: Akron 1158067153

117

much of the input energy is expended in dissipative processes, i.e., molecular motions.

For covalently crosslinked rubbers, the networks will be more elastic with increasing

crosslink densiy. Not much energy is dissipated. This leads to rupture at low strains.

Furthermore, covalent crosslinks cannot relax by breaking and reforming. When the

network chains break, molecular flaws will be created. If stresses around the molecular

flaws are magnified, catastrophic failure results.38 Therefore, the difference between

tensile strengths of ionically and covalently crosslinked rubbers is mainly due to the

ability to relax stress in the former case.

Figure 4.43 shows the breaking strains of XNBR cured by various curing agents.

Ultimate strains of covalently crosslinked rubbers decrease continuously with increased

crosslink density. For ionically crosslinked rubbers, breaking strains decrease at first, and

then seem to saturate. However, in the range of crosslink densities studied breaking

strains of ionically crosslinked rubbers are lower than those of covalently crosslinked

samples. This is due to a different topology of the networks. Aggregation of ionic

crosslinks will result in highly entangled networks, which will be broken at low strain.

4.4 ATR-IR Spectroscopy

4.4.1 Thermally Cured XNBR

ATR-IR spectra of uncured and thermally cured XNBR are shown in Figure 4.44.

A portion (1550 to 1850 cm-1) of these spectra is given in Figure 4.45. Characteristic of

the neat XNBR are peaks at 1697 cm-1 (carbonyl of acid dimer), 1730 cm-1 (carbonyl of

acid monomer), and 2235 cm-1 (triple bond of nitrile).110-113 The other designated peaks at

Page 136: Akron 1158067153

118

0.00 0.04 0.08 0.12 0.160

4

8

12

16

20

24 XN-MgA XN-MgB XN-MgC XN-P XN-Ch XN-Ca XN-Ba Heated XN

300%

Mod

ulus

(MPa

)

Vr

Ionic crosslinks

Covalent crosslinks

Figure 4.41 300% Modulus of the XNBR vulcanized by different curing agents as a function of Vr.

Page 137: Akron 1158067153

119

0.00 0.04 0.08 0.12 0.16 0.20

2

3

4

56789

10

20

30

40

50607080

Ten

sile

stre

ngth

(MPa

)

Vr

Ionic crosslinks XN-MgA XN-MgB XN-MgC XN-P XN-Ch XN-Ca XN-Ba Heated XN

Covalent crosslinks

Figure 4.42 Tensile strength of the XNBR vulcanized by different curing agents as a function of Vr.

Page 138: Akron 1158067153

120

0.00 0.04 0.08 0.12 0.16200

400

600

800

1000

1200

1400

1600 XN-MgA XN-MgB XN-MgC XN-P XN-Ch XN-Ca XN-Ba Heated XN

Elo

ngat

ion

at b

reak

(%)

Vr

Covalent crosslinks

Ionic crosslinks

Figure 4.43 Elongation at break of the XNBR vulcanized by different curing agents as a function of Vr.

Page 139: Akron 1158067153

121

920, 965, 1440, 1640, 1670, 2845, and 2920 cm-1 are contributed by hydrocarbon parts of

the elastomer backbone (Table 4.8). Evidently, most of carboxyl groups exist as the

hydrogen-bonded acid dimer. The corresponding O – H stretching frequency of the acid

dimer appears as a broad band at about 3200 cm-1 lying under the sharp stretching band

of C – H groups.105 Upon heating, the positions of the characteristic peaks do not change.

Qualitatively, the spectra of the cured samples are approximately the same as that of

unheated XNBR. However, in all cured samples the appearance of a small shoulder at

1750 to 1775 cm-1 is observed. Its origin is possibly due to the presence of anhydride

structure. Because carboxyl groups are randomly incorporated into the polymer backbone,

a butyric anhydride structure is more favorable than a cyclic structure (eq. 26). Grant and

Grassie106 studied thermal decomposition of poly(methacrylic acid) and reported that

characteristic butyric anhydride structures are shown by twin peaks at 1743 cm-1 and

1803 cm-1. Lee and coworkers105 studied the effect of temperature on anhydride

formation in poly(ethylene-co-methacrylic acid) and assigned characteristic frequencies

to butyric anhydride at 1735 cm-1 and 1780 cm-1. However, these bands appear as

shoulders in the spectra, not as the twin peaks observed by Grant and Grassie, probably

due to a difference in the content of methacrylic acid of studied polymers.

Page 140: Akron 1158067153

122

CH2

CH3C

CH2

CO

O

C

O

C CH3

CH2

CH2

Isobutyric anhydride

CH2

CH3C

CH2

C OH

O

C

O

C CH3

CH2

CH2

HO+ (26)

+H2O

Absorbance of the small shoulder in heated XNBR changes little with increased

heating time. This may be due an equilibrium which limits the amount of the anhydride

structure formed.106

Table 4.8 Characteristic group frequencies of the raw XNBR110-113

Wave number (cm-1) Assignment

920 Out-of-plane vibration of the methylene hydrogen atom of the vinyl group

965 Out-of-plane vibration of the hydrogen atom of the 1,4-trans component

1440 In-plane deformation of methylene group 1640 – 1670 Stretching of C = C

1697 Carbonyl stretching of hydrogen-bonded acid dimer 1730 Carbonyl stretching of monocarboxylic acid 2235 Stretching of nitrile triple bonds 2845 Symmetric stretching of methylene group 2920 Asymmetric stretching of methylene group

Page 141: Akron 1158067153

123

Figure 4.44 ATR-IR spectra of uncured and thermally cured XNBR in the range 800 to 4000 cm-1.

Page 142: Akron 1158067153

124

1850 1800 1750 1700 1650 1600 15500.0

0.1

0.2

0.3

0.4

0.5

0.6

1670

1640

1803

1730

Abs

orba

nce

Wave number (cm-1)

uncured cured 120 min cured 240 min cured 500 min cured 1000 min

Thermally cured raw XNBR at 165 oC

C

O

OH

C

O

O H

C

O

OH

1697

Anhydride crosslink

Figure 4.45 ATR-IR spectra of uncured and thermally cured XNBR in the range 1550 to 1850 cm-1.

Page 143: Akron 1158067153

125

4.4.2 XNBR-MgO Compositions

Figure 4.46 gives the ATR-IR spectra in the range 800 to 4000 cm-1 for uncured

neat XNBR and for uncured compositions containing high surface area MgO at 0.5x to

5.0x stoichiometry, assuming that one mole of MgO reacts with two moles of carboxyl

groups. Results in the range 1200 to 2000 cm-1 are shown in Figure 4.47. The peaks at

1697 cm-1 and 1730 cm-1 are assigned to carbonyl stretching in hydrogen-bonded and free

carboxylic acids, respectively. With the addition of MgO, intensities of these peaks

decrease, accompanied by a new peak at 1612 cm-1, together with the appearance of a

broad band centered at 3420 cm-1, which is assigned to vibration of OH groups (Table

4.9). Therefore, a peak at 1612 cm-1 is assigned to the magnesium hydroxycarboxylate,

– COOMgOH. A similar structure, zinc hydroxycarboxylate (– COOZnOH), which is a

product of reaction between carboxyl terminated polyester and ZnO, was assigned in the

same way.114 Carboxyl groups are essentially neutralized at stoichiometry of MgO, and

disappear when an equimolar amount of MgO was added. These results suggest that

neutralization occurs during mixing and continues during storage (samples were about 2

weeks old before collecting spectra).

The ATR-IR spectra of cured samples in the range 800 to 4000 cm-1 are given in

Figure 4.48. A portion (1200 to 2000 cm-1) of these spectra is shown in Figure 4.49. The

IR spectrum of cured XN-MgA0.5 is approximately the same as that of the uncured

sample. At stoichiometry (XN-MgA1.0), most of carboxyl groups are essentially

neutralized, and the salt peak becomes broader. With increased amounts (2.0x and 5.0x

stoichiometry) of MgO, a peak at 1587 cm-1 appears. This peak is attributed to

asymmetric carbonyl stretching of magnesium carboxylate salt (Table 4.9).66, 111 In the

Page 144: Akron 1158067153

126

XN-MgA0.5 vulcanizate, the magnesium hydroxycarboxylate (the peak at 1612 cm-1) is

the main product from neutralization as in the uncured rubber. This salt will not yield

efficient crosslinks and effective network chains. Although the cured XN-MgA0.5 has

higher tensile strength than the cured neat XNBR, it has lower Vr in the swollen gel and

higher sol content. A mix of ionic salts is obtained for the XN-MgA1.0 sample.

Apparently, a large amount of the magnesium carboxylate salt is obtained when at least

2.0x stoichiometry of MgO is present. This salt will give efficient crosslinks and effective

network chains. ATR-IR results are in good agreement with ODR, swelling and tensile

results, which show that equimolar amounts of MgO and carboxyl groups are needed to

give optimum properties. Therefore, the assumption that one mole of MgO reacts with

two carboxyl groups is incorrect.

Table 4.9 Characteristic group frequencies of XNBR-MgO compositions110-113

Wave number (cm-1) Assignment

920 Out-of-plane vibration of the methylene hydrogen atom of the vinyl group

965 Out-of-plane vibration of the hydrogen atom of the 1,4-trans component

1440 In-plane deformation of methylene group 1612 Carbonyl stretching of magnesium hydroxycarboxylate salt

1587 Asymmetric carbonyl stretching of magnesium carboxylate salt

1640 – 1670 Stretching of C = C 1697 Carbonyl stretching of hydrogen-bonded acid dimer 1730 Carbonyl stretching of monocarboxylic acid 2235 Stretching of nitrile triple bonds 2845 Symmetric stretching of methylene group 2920 Asymmetric stretching of methylene group 3420 Stretching of O – H

Page 145: Akron 1158067153

127

The proposed neutralizations (equations 22 to 24) are reasonably well-explained

by the behavior of the XNBR-MgO systems. Assume that all carboxyl groups have the

same reactivity. When MgO is added to the rubber, all carboxyl groups have equal

opportunity to react with MgO. But not all carboxylic acid groups are expected to react at

the same time, because of several reasons, for example: i) Because reactions occur in

solid state, and MgO particles are not dissolved in the rubber matrix, reactions are

expected to occur first at the surface of the MgO particles. The issues of surface area (or

particle size) and concentrations will then become important. That is cure reaction will be

controlled by the diffusion of the curing sites, ii) The diffusion of carboxyl groups can be

limited by the topology of the rubber matrix, i.e., by highly entangled rubber chains.

However, many carboxylic acid groups will react with MgO at the same time. The second

carboxyl group will not wait until the first one has reacted to form the magnesium

hydroxycarboxylate salt. Therefore, most of carboxylic acid groups will form the

magnesium hydroxycarboxylate salt. This type of salt does not give efficient crosslinks

and effective network chains, which is evidenced by a high degree of swelling, and a high

sol content. It is the coupling of magnesium hydroxycarboxylate salts to form magnesium

carboxylate that yields efficient crosslinks and effective network chains.

Page 146: Akron 1158067153

128

Figure 4.46 ATR-IR spectra of the uncured neat XNBR and XNBR-MgO compounds in the range 800 to 4000 cm-1.

Page 147: Akron 1158067153

129

2000 1800 1600 1400 12000.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

Abs

orba

nce

Wave number (cm-1)

Uncured XN-MgA compounds C

O

O H

C

O

OH

1697

C

O

OH

1730C

O

O MgOH

1612

1440

XNBR

XN-MgA0.5

XN-MgA1.0

XN-MgA2.0

XN-MgA5.0

Figure 4.47 ATR-IR spectra of the uncured neat XNBR and XNBR-MgO compounds in the range 1200 to 2000 cm-1.

Page 148: Akron 1158067153

130

Figure 4.48 ATR-IR spectra of the neat XNBR and XNBR-MgO vulcanizates in the range 800 to 4000 cm-1 (cured 120 min at 165 oC).

Page 149: Akron 1158067153

131

2000 1800 1600 1400 12000.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

C

O

O

C

O

O+ Mg +

C

O

O MgOH

C

O

OH

C

O

O H

C

O

OH

XN-MgO vulcanizates cured 120 min at 165 o CA

bsor

banc

e

Wave number (cm-1)

XNBR

XN-MgA0.5

XN-MgA1.0

XN-MgA2.0

XN-MgA5.0

1730

1697

1612

1587

1440

Figure 4.49 ATR-IR spectra of the neat XNBR and XNBR-MgO vulcanizates in the range 1200 to 2000 cm-1 (cured 120 min at 165 oC).

Page 150: Akron 1158067153

132

4.4.3 XNBR-CaO Compositions

ATR-IR spectra of uncured raw XNBR and XNBR-CaO compounds are given in

Figures 4.50. A portion (1550 to 1850 cm-1) of these spectra is shown in Figure 4.51. The

spectra of XN-Ca compounds are very similar to that of the neat XNBR. Unlike the

uncured XN-Mg compounds, no salt formation is observed in uncured XN-Ca samples.

They are the same as that of the neat XNBR (Table 4.8).

Figure 4.52 gives ATR-FTIR spectra of the neat XNBR and XNBR-CaO

vulcanizates. Spectra of XN-Ca vulcanizates are similar to that of the cured neat XNBR,

indicating similar cure mechanisms. Figure 4.53 shows spectra in the region 1550 to 1850

cm-1. A small shoulder in the range 1750 to 1775 cm-1 appears in all cured samples. The

difference between uncured and cured samples in the region 1550 to 1850 cm-1 is given

in Figure 4.54. The small shoulder is absent in all uncured compounds and appears in all

the cured samples. It may be attributed to anhydride structure. However, the absorption

frequencies of anhydride structures reported in the literature vary dependent on the

polymer.105, 106 Weak absorption in XN-Ca compositions may be due to an equilibrium in

condensate anhydride formation.106 As discussed earlier, swelling and tensile behavior of

XN-Ca vulcanizates are similar to those of the cured neat XNBR. This is strongly

supported by the ATR-IR results. Why these particular XNBR-CaO compositions do not

form ionic crosslinks is not clearly understood at this point.

Page 151: Akron 1158067153

133

Figure 4.50 ATR-IR spectra of the uncured neat XNBR and XNBR-CaO compounds in the range 800 to 4000 cm-1.

Page 152: Akron 1158067153

134

1850 1800 1750 1700 1650 1600 15500.0

0.1

0.2

0.3

0.4

0.5

0.6A

bsor

banc

e

Wave number (cm-1)

Uncured XN-CaO compounds XNBR XN-Ca0.5 XN-Ca1.0 XN-Ca2.0 XN-Ca5.0

C

O

O H

C

O

OH

1697

C

O

OH

1730

Figure 4.51 ATR-IR spectra of the uncured neat XNBR and XNBR-CaO compounds in the range 1550 to 1850 cm-1.

Page 153: Akron 1158067153

135

Figure 4.52 ATR-IR spectra of the neat XNBR and XNBR-CaO vulcanizates in the range 800 to 4000 cm-1 (cured 1000 min at 165 oC).

Page 154: Akron 1158067153

136

1850 1800 1750 1700 1650 1600 15500.0

0.1

0.2

0.3

0.4

0.5A

bsor

banc

e

Wave number (cm-1)

XN-CaO vulcanizates cured 1000 min at 165 oC

XNBR XN-Ca0.5 XN-Ca1.0 XN-Ca2.0 XN-Ca5.0

C

O

O H

C

O

OH

1697

C

O

OH

1732

Anhydridecrosslink

Figure 4.53 ATR-IR spectra of the neat XNBR and XNBR-CaO vulcanizates in the range 1550 to 1850 cm-1 (cured 1000 min at 165 oC).

Page 155: Akron 1158067153

137

1850 1800 1750 1700 1650 1600 15500.0

0.1

0.2

0.3

0.4

0.5A

bsor

banc

e

Wave number (cm-1)

C

O

O H

C

O

OH

1697

C

O

OH

1730

Uncured XNBR Cured XNBR Uncured XN-Ca2.0 Cured XN-Ca2.0

Anhydride crosslink

Cured 1000 min at 165 oC

Figure 4.54 ATR-IR spectra of the uncured neat XNBR and XN-Ca2.0 compounds, and the neat XNBR and XN-Ca2.0 vulcanizates in the range 1550 to 1850 cm-1 (cured 1000 min at 165 oC).

Page 156: Akron 1158067153

138

4.4.4 XNBR-Ca(OH)2 Compositions

ATR-FTIR spectra of the uncured neat XNBR and XNBR-Ca(OH)2 compounds

are given in Figure 4.55. Absorption of these samples in the range 1200 to 2000 cm-1 is

shown in Figure 4.56. The spectra are characterized by the peaks at 1697 cm-1 (acid

dimers), 1730 cm-1 (a free acid), and 2235 cm-1 (nitrile). The difference between spectra

of XNBR-Ca(OH)2 compounds from that of the neat XNBR is the presence of a peak at

3642 cm-1, which becomes prominent with increased amounts of Ca(OH)2 (Table 4.10).

This peak is contributed to vibration of Ca(OH)2. No peaks are observed in the frequency

region (1500 to 1600 cm-1) of asymmetric stretching of carboxylate anion, suggesting that

neutralization does not occur before curing.

Upon curing, the spectrum of the neat XNBR remains largely the same as that of

the uncured sample, except for the appearance of a small shoulder in the range 1750 to

1775 cm-1, which may be attributed to the anhydride structure type. However, large

amounts of carboxylic acid groups remain unreacted, evidenced by strong intensities of

the peaks at 1697 cm-1 (acid dimer), and 1730 (acid monomer), respectively. For the

compounds containing Ca(OH)2, the intensities of the acid peaks decrease, accompanied

by a new peak at 1560 cm-1 (Figures 4.57 and 4.58). This peak is assigned to asymmetric

carbonyl stretching of calcium carboxylate.65, 66, 111, 115 In a XN-Ch0.5 vulcanizate, there

is a certain amount of unreacted carboxyl groups left in the sample, and the salt peak is

broad. Carboxyl groups are essentially neutralized at a stoichiometric amount of Ca(OH)2.

Complete neutralization is obtained when at least 2.0x stoichiometry of Ca(OH)2 is

present, and the salt peak becomes narrower.

Page 157: Akron 1158067153

139

Table 4.10 Characteristic group frequencies of XNBR-Ca(OH)2 samples65, 66, 110, 111, 113

Wave number (cm-1) Assignment

920 Out-of-plane vibration of the methylene hydrogen atom of the vinyl group

965 Out-of-plane vibration of the hydrogen atom of the 1,4-trans component

1410 Symmetric carbonyl stretching of calcium carboxylate 1440 In-plane deformation of methylene group 1560 Asymmetric carbonyl stretching of calcium carboxylate salt

1640 – 1670 Stretching of C = C 1697 Carbonyl stretching of hydrogen-bonded acid dimer 1730 Carbonyl stretching of monocarboxylic acid 2235 Stretching of nitrile triple bonds 2845 Symmetric stretching of methylene group 2920 Asymmetric stretching of methylene group 3400 Stretching of O – H 3642 Ca(OH)2

It seems that the structure of the ionic salts formed in XN-Ch1.0, XN-Ch2.0, and

XN-Ch5.0 is very similar, but the amount of salt formed is different. In fact, these three

samples show a substantial increase in tensile strength compared to the neat XNBR, 42

MPa for XN-Ch1.0, 51 MPa for XN-Ch2.0, and 48 MPa for XN-Ch5.0. In compositions

containing up to 2.0x stoichiometry, Ca(OH)2 is completely used in neutralization of

carboxyl groups, evidenced by disappearance of the Ca(OH)2 peak at 3642 cm-1. In the

XN-Ch5.0 vulcanizate, however, unreacted Ca(OH)2 is observed. Apparently, the excess

amount of Ca(OH)2 has a little effect on the tensile properties of the vulcanizates.

Page 158: Akron 1158067153

140

Figure 4.55 ATR-IR spectra of the uncured neat XNBR and XNBR-Ca(OH)2 compounds in the range 800 to 4000 cm-1.

Page 159: Akron 1158067153

141

2000 1800 1600 1400 1200

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

Abs

orba

nce

Wave number (cm-1)

Uncured XN-Ca(OH)2 compounds

C

O

O H

C

O

OH

C

O

OH

1730

1697 1440

XNBR

XN-Ch0.5

XN-Ch1.0

XN-Ch2.0

XN-Ch5.0

Figure 4.56 ATR-IR spectra of the uncured neat XNBR and XNBR-Ca(OH)2 compounds in the range 1200 to 2000 cm-1.

Page 160: Akron 1158067153

142

Figure 4.57 ATR-IR spectra of the neat XNBR and XNBR-Ca(OH)2 vulcanizates in the range 800 to 4000 cm-1.

Page 161: Akron 1158067153

143

2000 1800 1600 1400 1200

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

1410

XN-Ca(OH)2 vulcanizates cured 240 min at 165 oC

1440

Abs

orba

nce

Wave number (cm-1)

C

O

O H

C

O

OH

1697C

O

OH

17301560

C

O

O

C

O

O+ Ca +

XNBR

XN-Ch0.5

XN-Ch1.0

XN-Ch2.0

XN-Ch5.0

Figure 4.58 ATR-IR spectra of the neat XNBR and XNBR-Ca(OH)2 vulcanizates in the range 1200 to 2000 cm-1.

Page 162: Akron 1158067153

144

4.4.5 XNBR-BaO Compositions

ATR-FTIR spectra of the uncured raw XNBR and XNBR-BaO compounds are

shown in Figures 4.59 (in the range 800 to 4000 cm-1) and 4.60 (in the range 1200 to

2000 cm-1), respectively. Characteristic group frequencies of these spectra are listed in

Table 4.11. Apparently, ATR-IR spectra of uncured XN-Ba compounds are

approximately the same as that of the uncured neat XNBR, characterized by the peaks at

1697 cm-1 (acid dimer), 1730 cm-1 (acid monomer), and 2235 cm-1 (nitrile). No

absorption in the region 1500 to 1600 cm-1 is observed, indicating that no salt structure

has developed before curing.

Figure 4.61 shows the ATR-IR spectra in the region 800 to 4000 cm-1 of the neat

XNBR and XNBR-BaO vulcanizates. A portion (1200 to 2000 cm-1) of these spectra is

given in Figure 4.62. Frequencies of characteristic functional groups are assigned in

Table 4.11, and some of them are marked in the Figures. As discussed earlier, the neat

XNBR undergoes self-crosslinking upon curing. However, large amounts of carboxyl

groups are expected to remain, which is evidenced by strong absorption of the dimeric

acid (1697 cm-1), and free acid (1730 cm-1), respectively.

Upon heating XN-Ba compounds, carboxyl groups are neutralized to form ionic

salts. This is evidenced by the decrease in intensities of the acid peaks (1697 and 1730

cm-1) accompanied by appearance of the peak at 1546 cm-1, assigned to asymmetric

carbonyl stretching of barium carboxylate salts. Degree of neutralization increases with

the increased amounts of BaO. Although carboxyl groups are mainy neutralized at 2.0x

stoichiometry of BaO, there are still unreacted carboxyl groups. This is due to BaO that is

not dissolved and well-dispersed in the XNBR matrix. The poor dispersion and large

Page 163: Akron 1158067153

145

particle size also account for the inferior tensile properties of XN-Ba vulcanizates

compared to XNBR-MgO and XNBR-Ca(OH)2 vulcanizates. However, complete

neutralization is observed in the XN-Ba5.0 vulcanizate.

Table 4.11 Characteristic group frequencies of XNBR-BaO samples

Wave number (cm-1) Assignment

920 Out-of-plane vibration of the methylene hydrogen atom of the vinyl group

965 Out-of-plane vibration of the hydrogen atom of the 1,4-trans component

1405 Symmetric carbonyl stretching of barium carboxylate 1440 In-plane deformation of methylene group 1546 Asymmetric carbonyl stretching of calcium carboxylate salt

1640 – 1670 Stretching of C = C 1697 Carbonyl stretching of hydrogen-bonded acid dimer 1730 Carbonyl stretching of monocarboxylic acid 2235 Stretching of nitrile triple bonds 2845 Symmetric stretching of methylene group 2920 Asymmetric stretching of methylene group 3400 Stretching of O – H

4.4.6 Comparison among Metal Compounds

Figure 4.63 shows ATR-IR spectra in the range 1475 to 1850 cm-1 for the cured

neat XNBR and XNBR vulcanized by 2.0x stoichiometry of metal compounds. The

spectra of the neat XNBR and XN-Ca2.0 are essentially the same. No peaks are observed

in the frequency region (1500 to 1600 cm-1) for asymmetric carbonyl stretching of

carboxylate groups. Curing mechanisms of these two samples involve coupling

Page 164: Akron 1158067153

146

Figure 4.59 ATR-IR spectra of the uncured neat XNBR and XNBR-BaO compounds in the range 800 to 4000 cm-1.

Page 165: Akron 1158067153

147

2000 1800 1600 1400 1200

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

Abs

orba

nce

Wave number (cm-1)

Uncured XN-Ba compounds

C

O

O H

C

O

OH

1697C

O

OH

1730

1440

XNBR

XN-Ba0.5

XN-Ba1.0

XN-Ba2.0

XN-Ba5.0

Figure 4.60 ATR-IR spectra of the uncured neat XNBR and XNBR-BaO compounds in the range 1200 to 2000 cm-1.

Page 166: Akron 1158067153

148

Figure 4.61 ATR-IR spectra of the neat XNBR and XNBR-BaO vulcanizates in the range 800 to 4000 cm-1.

Page 167: Akron 1158067153

149

2000 1800 1600 1400 1200

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8XN-BaO vulcanizates cured 240 min at 165 oC

Abs

orba

nce

Wave number (cm-1)

C

O

O H

C

O

OH

C

O

O

C

O

O+ Ba +

C

O

OH 1697

17301546

1440

1405

XNBR

XN-Ba0.5

XN-Ba1.0

XN-Ba2.0

XN-Ba5.0

Figure 4.62 ATR-IR spectra of the neat XNBR and XNBR-BaO vulcanizates in the range 1200 to 2000 cm-1.

Page 168: Akron 1158067153

150

of carboxyl groups to form anhydride crosslinks. In contrast to the neat XNBR and XN-

Ca2.0 vulcanizates, ionic crosslinks are formed in XN-MgA2.0, XN-Ch2.0, and XN-

Ba2.0 vulcanizates, as evidenced by the disappearance of the acid peaks (1697 and 1730

cm-1) accompanied by new peaks corresponding to asymmetric carbonyl stretching of

carboxylate groups in the range 1500 to 1600 cm-1. The difference among ATR-IR

spectra of XN-MgA2.0, XN-Ch2.0, and XN-Ba2.0 vulcanizates is in the frequency region

of asymmetric carbonyl stretching. Apparently, these frequencies decrease in a

predictable manner with increasing cation mass and size. Similar results were obtained by

Brozoski and coworkers,66 and Han and Williams,115 who utilized infrared spectroscopy

to study local structures of copolymers of ethylene-methacrylic acid neutralized by alkali,

alkaline earth, and transition metals.

In the case of vibration of a simple harmonic oscillator, the relationship between

the wave number (ν′) of the absorption peak and the vibration frequency of bonds in the

molecule is given by

21

21

mm)mm(f

c21f

c21 +

π=

μπ=ν′ (26)

where f is the force constant of the bond (dyne/cm or g/s2), c is the velocity of light

(2.998 x 1010 cm/s), and m1, m2 are the masses (g) of atoms 1 and 2, respectively.110, 115

Assume that this principle can be applied in our case, and that metal ion types influence

the asymmetric stretching of carboxylate anions. Because Mg++ ion has the lowest mass

and size, and according to swelling results, Mg++ ion forms the strongest ionic bond with

the carboxylate anion when compared with Ca++ and Ba++ ions, therefore, the absorption

Page 169: Akron 1158067153

151

1850 1800 1750 1700 1650 1600 1550 1500

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

C

O

OH

C

O

O H

C

O

OH

C

O

O

C

O

O+ Ba +

C

O

O

C

O

O+ Ca +

C

O

O

C

O

O+ Mg +

Abs

orba

nce

Wave number (cm-1)

XNBR XN-MgA2.0 XN-Ca2.0 XN-Ch2.0 XN-Ba2.0

1697

1730

1587

1560

1546

Anhydridecrosslink

Figure 4.63 ATR-IR spectra of the neat XNBR, XN-MgA2.0, XN-Ca2.0, XN-Ch2.0, and XN-Ba2.0 vulcanizates in the range 1475 to 1850 cm-1.

Page 170: Akron 1158067153

152

by asymmetric stretching by the magnesium carboxylate is expected to be at a higher

wave number than for Ca++ and Ba++ ions. Although this assumption is oversimplified, it

explains the shift in absorption frequency of asymmetric carboxylate stretching with the

increasing mass and size of alkaline earth metal ions.

4.5 Dynamic Mechanical Properties

4.5.1 XNBR-MgO Vulcanizates

Substantial reinforcement of XNBR vulcanized by MgO is attributed to

aggregation of ionic crosslinks to form hard ionic domains which act as multifunctional

crosslinks and reinforcing structures.53, 56, 71 Dynamic mechanical analysis has commonly

been employed to study these ionic aggregates.59, 62, 64, 89-91, 93, 94, 116, 117 The storage

moduli (E′) of XN-MgA and XN-peroxide vulcanizates as a function of temperature at a

frequency 1.0 Hz are shown in Figure 4.64. Glassy moduli of all the vulcanizates are

similar. With increasing temperature passing through the transition zone, the moduli of

all samples drop sharply by about three orders of magnitude. This is the well-known

glass-rubber transition which arises from segmental relaxation of polymer chains. In the

rubbery zone, the storage moduli of XN-P1.0 and XN-MgA are quite different. The

modulus of XN-P1.0, which is covalently crosslinked, is lower than those of ionically

crosslinked XN-MgA vulcanizates, but it does not change with increasing temperature,

suggesting stability of crosslinks towards heat. XN-MgA vulcanizates have greater

moduli due to aggregation of ionic species to form hard domains. The rubbery modulus

increases with increased amount of MgO from 1.0x (XN-MgA1.0) to 2.0x (XN-MgA2.0)

stoichiometry, and slightly thereafter (XN-MgA3.0 sample). This is due to an increase in

Page 171: Akron 1158067153

153

the number of effective ionic crosslinks, from the magnesium carboxylate salt. Because

of the physical nature of ionic crosslinks, the rubbery modulus decreases with increasing

temperature. This is not surprising because in the XN-MgA1.0 sample most of the ionic

salt is magnesium hydroxycarboxylate (eq. 22), which is formed intramolecularly.

Because of the polar nature of this salt, it yields ineffective ionic crosslinks. The ionic

crosslinks are expected to be thermally labile, as shown by the decrease in rubbery

modulus with increasing temperature. In the vulcanizates containing MgO 2.0x

stoichiometry or more, the magnesium carboxylate salt (eq. 23) is the main product

which links two polymer chains together. This salt gives effective crosslinks, which are

expected to be less thermally labile. Therefore, the rubbery modulus is more stable

towards heat.

The dynamic loss moduli (E″) of XN-P and XN-MgA vulcanizates as a function

of temperature are given in Figure 4.65. In the glassy zone, the loss moduli of all

vulcanizates are similar, suggesting that input energy is dissipated in similar processes.

However, the behavior of XN-P1.0 and XN-MgA vulcanizates in the rubbery region is

different. The covalently crosslinked XN-P1.0 has the lowest dynamic loss modulus,

suggest that energy dissipation is much smaller than for XN-MgA vulcanizates.

Apparently, there are other loss processes at high temperatures (20 to 100 oC).

Surprisingly, the loss modulus becomes larger with increased amounts of MgO, although

the elastic modulus rises (Figure 4.64). The dissipative processes have been suggested to

involve ion hopping processes or migration of ion pairs attached to a particular polymer

chain segment from one ionic aggregate to another (Figure 4.66).118-120 This concept is

Page 172: Akron 1158067153

154

-100 -50 0 50 100 150

106

107

108

109

XN-MgA vulcanizates cured 120 min at 165 oC

XN-P1.0

XN-MgA2.0 XN-MgA3.0

E' (

Pa)

Temperature (oC)

XN-MgA1.0

Figure 4.64 Temperature dependence of dynamic storage modulus (E′) of XN-P1.0 and XN-MgA vulcanizates at frequency 1.0 Hz.

Page 173: Akron 1158067153

155

-100 -50 0 50 100 150

105

106

107

108

XN-MgA vulcanizates cured 120 min at 165 oC

XN-P1.0

XN-MgA2.0XN-MgA3.0

E"

(Pa)

Temperature (oC)

XN-MgA1.0

Figure 4.65 Temperature dependence of dynamic loss modulus (E″) of XN-P1.0 and XN-MgA vulcanizates at frequency 1.0 Hz.

Page 174: Akron 1158067153

156

Figure 4.66 Schematic drawing of ion hopping mechanisms (opened and closed circles represent ion pairs).121

Page 175: Akron 1158067153

157

closely related to bond interchange which has been used to explain the strength of these

elastomers.51

The temperature dependence of loss tangent (tan δ) of XN-P1.0 and XN-MgA

vulcanizates is shown in Figure 4.67. Molecular transitions corresponding to peaks

appearing in the plot are listed in Table 4.12. In the XN-P1.0 sample only the glass

transition is observed. For the XN-MgA vulcanizates, there is another broad transition at

a high temperature other than the glass transition of the rubber matrix. This transition is

not observed in the XN-P1.0 sample, where the rubber molecules are covalently

crosslinked. Therefore, this transition must involve the ionic aggregates formed in the

XN-MgA vulcanizates. Its origin has been suggested to arise from relaxation processes

by exchange of ion pairs between ionic aggregates (Figure 4.66).91, 119 However, the

precise position of the ionic transition is difficult to determine due to lack of a clear

maximum. For XNBR vulcanized by zinc oxide, a similar peak in the range 50 to 80 oC

has been reported, depending on the testing frequency.62, 64, 89, 90, 116 With increasing MgO

content, the transition shifts to higher temperatures. Similar observations has been

reported for the ZnO-cured XNBR90 and ZnO-activated sulfur-cured XSBR.59

The effect of the specific surface area of MgO on the storage modulus of XNBR

vulcanizates is shown in Figure 4.68. The glassy moduli of all vulcanizates are similar.

However the high surface area (type A) MgO gives vulcanizates with a slightly higher

rubber modulus than does the low surface area MgO (type C). This is due to the

difference in the amount of effective ionic crosslinks formed (Table 4.2). As in the case

of XN-MgA, the rubbery moduli of XN-MgC increase with increased concentration of C-

type MgO, and the rubbery modulus becomes more thermally stable.

Page 176: Akron 1158067153

158

-100 -50 0 50 100 150

0.02

0.03

0.040.050.060.070.080.090.1

0.2

0.3

0.40.50.60.70.80.9

1

XN-MgA vulcanizates cured 120 min at 165 oC

XN-MgA3.0

XN-MgA2.0

tan

δ

Temperature (oC)

XN-P1.0

XN-MgA1.0

Figure 4.67 Temperature dependence of loss tangent (tan δ) of XN-P1.0 and XN-MgA vulcanizates at frequency 1.0 Hz.

Page 177: Akron 1158067153

159

Table 4.12 Molecular transition temperatures of XN-P1.0, and XN-MgA vulcanizates at a frequency of 1.0 Hz

Vulcanizates Tg (oC)* Ionic transition temperature range (oC)

XN-P1.0 -24 Not observed

XN-MgA1.0 -22 10 - 100

XN-MgA2.0 -20 20 - 110

XN-MgA3.0 -18 25 - 115 * Taken from the temperature at a tan δ maximum

Figure 4.69 shows the dependence of the dynamic loss modulus of XN-Mg

vulcanizates on the specific surface area of MgO. The loss moduli in the glassy zone of

all vulcanizates are similar. As in the case of XN-MgA vulcanizates, a small hump in the

rubbery region appears in XN-MgC vulcanizates, and becomes more pronounced with

increasing amounts of MgO. As discussed earlier, the origin of the small hump, i.e. more

energy lost, is believed to be due to the migration of ion pairs attached to a particular

polymer chain from one ionic aggregate to another.91, 118-120 At the same concentration,

high surface area MgO (type A) results in a slightly higher loss modulus than for the low

surface area MgO (type C). This is probably due to a difference in the amount of ionic

crosslinks formed.

The effect of specific surface area of MgO on the loss tangent (tan δ) of XN-Mg

vulcanizates is illustrated in Figure 4.70. Two transitions are observed, as for XN-MgA.

The transition at low temperature is the glass transition of the rubber matrix. Another

transition at high temperature is associated with ionic species, because it is absent in

Page 178: Akron 1158067153

160

-100 -50 0 50 100 150

106

107

108

109

XN-MgC2.0

XN-MgA1.0

XN-MgC1.0

XN-MgA2.0

Frequency 1.0 Hz XN-MgA1.0 XN-MgC1.0 XN-MgA2.0 XN-MgC2.0 XN-P1.0

A = 140 m2/gC = 45 m2/g

E' (

Pa)

Temperature (oC)

XN-P1.0

Figure 4.68 Effect of specific surface area on dynamic storage modulus (E′) of XN-Mg vulcanizates at frequency 1.0 Hz.

Page 179: Akron 1158067153

161

-100 -50 0 50 100 150

105

106

107

108

XN-MgA1.0 XN-MgC1.0 XN-MgA2.0 XN-MgC2.0 XN-P1.0

A = 140 m2/gC = 45 m2/g

E"

(Pa)

Temperature (oC)

Frequency 1.0 Hz

XN-MgA2.0

XN-MgC2.0

XN-MgA1.0

XN-MgC1.0

XN-P1.0

Figure 4.69 Effect of specific surface area on dynamic loss modulus (E″) of XN-Mg vulcanizates at frequency 1.0 Hz.

Page 180: Akron 1158067153

162

-100 -50 0 50 100 150

0.1

1

XN-MgA1.0 XN-MgC1.0 XN-MgA2.0 XN-MgC2.0 XN-P1.0

A = 140 m2/gC = 45 m2/g

tan

δ

Temperature (oC)

Frequency 1.0 Hz

XN-MgA2.0

XN-MgC2.0

XN-MgA1.0XN-MgC1.0

XN-P1.0

Figure 4.70 Effect of specific surface area on loss tangent (tan δ) of XN-Mg vulcanizates at frequency 1.0 Hz.

Page 181: Akron 1158067153

163

covalently cured XNBR. The ionic transition is dependent on both the concentration and

the surface area of MgO. It shifts to higher temperatures with increase in both

concentration and surface area.

4.5.2 XNBR-CaO Vulcanizates

The temperature dependence of the dynamic storage modulus (E′) of XN-P1.0

and XNBR-CaO vulcanizates is given in Figure 4.71. All vulcanizates have similar glassy

and rubbery moduli. XN-P1.0 has a slightly higher rubbery modulus relative to CaO-

cured specimens. This is because curing in XNBR-CaO systems occurs via coupling

reaction of carboxyl groups to form anhydride crosslinks, which are limited by the

amount of carboxyl groups. However, all vulcanizates behave similarly. The rubbery

zone in XN-Ca vulcanizates is almost a plateau, indicating the stability of crosslinks

towards heat. This is indirect evidence that most of crosslinks in XN-Ca vulcanizates are

covalent.

The effect of temperature on the dynamic loss modulus (E″) of XN-P1.0 and

XNBR-CaO samples is also very similar (Figure 4.72). XN-Ca vulcanizates have slightly

higher loss modulus than XN-P1.0 in the rubbery region. No peak from ionic aggregates

is observed.

Figure 4.73 shows a plot of tan δ of XN-P1.0 and XN-Ca vulcanizates against

temperature. Only one peak at a temperature of about -22 to -24 oC is observed, which is

the glass transition temperature of the XNBR matrix (Table 4.13). In contrast to XN-Mg

vulcanizates, no ionic transition is observed. No ionic aggregates are formed in these

systems.

Page 182: Akron 1158067153

164

-100 -50 0 50 100 150 200

106

107

108

109

XN-Ca3.0

XN-Ca2.0

E' (

Pa)

Temperature (oC)

XN-P1.0

XN-Ca1.0

XN-CaO vulcanizates cured 1000 min at 165 oC

Figure 4.71 Temperature dependence of dynamic storage modulus (E′) of XN-P1.0 and XN-Ca vulcanizates at frequency 1.0 Hz.

Page 183: Akron 1158067153

165

-100 -50 0 50 100 150 200

105

106

107

108

E"

(Pa)

Temperature (oC)

XN-CaO vulcanizates cured 1000 min at 165 oC

XN-P1.0

XN-Ca3.0XN-Ca2.0

XN-Ca1.0

Figure 4.72 Temperature dependence of dynamic loss modulus (E″) of XN-P1.0 and XN-Ca vulcanizates at frequency 1.0 Hz.

Page 184: Akron 1158067153

166

-100 -50 0 50 100 150 200

0.1

1 XN-P1.0 XN-Ca1.0 XN-Ca2.0 XN-Ca3.0

tan

δ

Temperature (oC)

XN-CaO vulcanizates cured 1000 min at 165 oC

Figure 4.73 Temperature dependence of loss tangent (tan δ) of XN-P1.0 and XN-Ca vulcanizates at frequency 1.0 Hz

Page 185: Akron 1158067153

167

Table 4.13 Molecular transition temperatures of XN-P1.0, and XN-CaO vulcanizates at frequency 1.0 Hz

Vulcanizates Tg (oC) Ionic transition temperature range (oC)

XN-P1.0 -24 Not observed

XN-Ca1.0 -24 Not observed

XN-Ca2.0 -22 Not observed

XN-Ca3.0 -22 Not observed

4.5.3 XNBR-Ca(OH)2 Vulcanizates

The dynamic storage modulus, loss modulus, and tan δ as a function of

temperature of XN-P1.0 and XNBR-Ca(OH)2 are given in Figures 4.74, 4.75, and 4.76,

respectively. No significant difference in the storage moduli in the glassy region is

observed. The glass transition slightly shifts to higher temperatures with increased

amount of Ca(OH)2. However, the behavior of covalently crosslinked (XN-P1.0) and

ionically crosslinked (XN-Ch) rubbers are quite different. The rubbery modulus of XN-

P1.0 is the lowest, but the most thermally stable, suggesting that the crosslinks are

permanent. At the beginning of the rubbery region, XN-Ch1.0 has a higher modulus than

XN-P1.0, but the modulus decreases with increased temperature, suggesting that the

crosslinks are unstable towards heat. At temperatures above 150 oC, the rubbery modulus

of XN-Ch1.0 becomes less than that of XN-P1.0. Although XN-Ch1.0 has high tensile

strength (~ 40 MPa) due to the ionic salt, calcium carboxylate, that is formed, ATR-IR

results show that not all carboxyl groups are neutralized. The decrease in the rubbery

Page 186: Akron 1158067153

168

modulus with temperature suggests that large ionic aggregates may not form at this

concentration of Ca(OH)2. With increasing Ca(OH)2 concentration, the rubbery modulus

increases greatly and is almost an order of magnitude higher than that of XN-P1.0. The

rubbery zone becomes more of a plateau due to the contribution from ionic aggregates.

At above 150 oC the storage modulus begins to drop, indicating thermal instability of the

crosslinks.

The loss moduli in the glassy region of all cured specimens are similar. However,

in the rubbery zone, the loss moduli of XN-Ch vulcanizates are greater than for XN-P1.0,

although XN-Ch vulcanizates have higher storage moduli. This is contributed to ionic

aggregates.

The plot (Figure 4.76) of tan δ against temperature of XN-Ch samples reveals two

transitions, the glass transition of the rubber matrix at low temperatures and the ionic

transition at high temperatures (Table 4.14). However, at high concentrations of Ca(OH)2

it is difficult to determine the precise position of the ionic transition because there is no

clear maximum. The mechanism of the ionic transition has been suggested to arise from

the interchange of ion pairs between ionic aggregates (Figure 4.66).91, 119-121 As in the

case of XN-Mg samples, the ionic transition of XN-Ch vulcanizates shifts to higher

temperatures. In contrast to XN-Ch vulcanizates, XN-P1.0 sample has only one transition,

corresponding to the glass transition of the rubber matrix.

Page 187: Akron 1158067153

169

-100 -50 0 50 100 150 200

106

107

108

109

XN-Ca(OH)2 vulcanizates cured 240 min at 165 oC

XN-Ch3.0XN-Ch2.0

XN-P1.0

E' (

Pa)

Temperature (oC)

XN-Ch1.0

Figure 4.74 Temperature dependence of dynamic storage modulus (E′) of XN-P1.0 and XN-Ch vulcanizates at frequency 1.0 Hz.

Page 188: Akron 1158067153

170

-100 -50 0 50 100 150 200

105

106

107

108

XN-Ca(OH)2 vulcanizates cured 240 min at 165 oC

E"

(Pa)

Temperature (oC)

XN-Ch2.0

XN-Ch3.0

XN-Ch1.0

XN-P1.0

Figure 4.75 Temperature dependence of dynamic loss modulus (E″) of XN-P1.0 and XN-Ch vulcanizates at frequency 1.0 Hz.

Page 189: Akron 1158067153

171

-100 -50 0 50 100 150 200

0.1

1

tan

δ

Temperature (oC)

XN-Ca(OH)2 vulcanizates cured 240 min at 165 oC

XN-Ch3.0

XN-Ch2.0

XN-Ch1.0

XN-P1.0

Figure 4.76 Temperature dependence of loss tangent (tan δ) of XN-P1.0 and XN-Ch vulcanizates at frequency 1.0 Hz.

Page 190: Akron 1158067153

172

Table 4.14 Molecular transition temperatures of XN-P1.0 and XN-Ca(OH)2 vulcanizates at frequency 1.0 Hz

Vulcanizates Tg (oC) Ionic transition temperature range (oC)

XN-P1.0 -24 Not observed

XN-Ch1.0 -18 20 - 145

XN-Ch2.0 -15 Not clear

XN-Ch3.0 -16 Not clear

4.5.4 XNBR-BaO Vulcanizates

The effect of temperature on dynamic storage modulus, loss modulus, and tan δ of

XN-P1.0 and XN-Ba vulcanizates is shown in Figures 4.77, 4.78, and 4.79, respectively.

As for XN-Mg and XN-Ch vulcanizates, the rubbery moduli of cured XN-Ba samples are

higher than for XN-P1.0 at the beginning of the rubbery zone, and increase with

increasing BaO concentrations due to the increase in number of ionic crosslinks. Because

of the physical nature of ionic crosslinks, the rubbery moduli of XN-Ba vulcanizates

decrease with temperature, and are finally lower than for XN-P1.0, indicating that the

ionic crosslinks are unstable towards heat. In the case of covalently crosslinked XN-P1.0,

the rubbery modulus remains constant over the entire range of temperature because

crosslinks are permanent. In the ionically crosslinked rubbers, the glass transition shifts

slightly to higher temperatures than for the covalently crosslinked rubber.

The dynamic loss moduli in the glassy zone are similar for all the vulcanizates

(Figure 4.78). The XN-Ba vulcanizates have a higher loss modulus than XN-P1.0

Page 191: Akron 1158067153

173

samples in the rubbery zone, and the loss moduli increase with the increase in BaO

concentrations. The increase in loss modulus is probably due to energy dissipated in

interchange of ion pairs.91, 119-121

The effect of temperature on tan δ of cured XN-P1.0 and XB-Ba samples is

shown in Figure 4.79. As in the case of XN-Mg and XN-Ch vulcanizates, two transitions

are observed in the cured XN-Ba specimens, the glass transition of the rubber matrix at a

low temperature, and the ionic transition at a high temperature (Table 4.15). However, it

is difficult to determine the precise position of the ionic transition due to there is no clear

maximum. For the covalently crosslinked sample only the glass transition is observed.

The glass transition shifts to higher temperature in the ionically crosslinked rubbers.

Table 4.15 Molecular transition temperatures of XN-P1.0 and XN-CaO vulcanizates at frequency 1.0 Hz

Vulcanizates Tg (oC) Ionic transition temperature range (oC)

XN-P1.0 -24 Not observed

XN-Ba1.0 -18 Not clear

XN-Ba2.0 -15 25 – 140

XN-Ba3.0 -18 25 – 140

Page 192: Akron 1158067153

174

-100 -50 0 50 100 150 200

106

107

108

109

XN-Ba3.0

XN-Ba2.0

XN-Ba1.0

E' (

Pa)

Temperature (oC)

XN-BaO vulcanizates cured 240 min at 165 oC

XN-P1.0

Figure 4.77 Temperature dependence of dynamic storage modulus (E′) of XN-P1.0 and XN-Ba vulcanizates at frequency 1.0 Hz.

Page 193: Akron 1158067153

175

-100 -50 0 50 100 150 200

105

106

107

108

XN-BaO vulcanizates cured 240 min at 165 oC

E"

(Pa)

Temperature (oC)

XN-Ba3.0

XN-Ba2.0

XN-Ba1.0

XN-P1.0

Figure 4.78 Temperature dependence of dynamic loss modulus (E″) of XN-P1.0 and XN-Ba vulcanizates at frequency 1.0 Hz

Page 194: Akron 1158067153

176

-100 -50 0 50 100 150 200

0.1

1

XN-BaO vulcanizates cured 240 min at 165 oCta

n δ

Temperature (oC)

XN-Ba3.0

XN-Ba2.0

XN-Ba1.0

XN-P1.0

Figure 4.79 Temperature dependence of loss tangent (tan δ) of XN-P1.0 and XN-Ba vulcanizates at frequency 1.0 Hz.

Page 195: Akron 1158067153

177

4.5.5 Comparison among Metal Compounds

Figure 4.80 shows tan δ as a function of temperature for peroxide-cured XNBR

and XNBR cured with 2.0x stoichiometry of various metal compounds. Clearly, two

transitions are observed in XN-MgA2.0, XN-Ch2.0 and XN-Ba2.0 vulcanizates. One is

the glass-rubber transition of the rubber matrix; the other, a high temperature is the ionic

transition. The ionic transition does not appear in XN-P1.0 and XN-Ca2.0 where the

rubber chains are covalently crosslinked. It appears only in the vulcanizates that are

crosslinked ionically, and may be associated with the exchange of ion pairs between ionic

aggregates. It seems to shift to higher temperature with increasing cation size.

Page 196: Akron 1158067153

178

-100 -50 0 50 100 150 200

0.1

1

tan

δ

Temperature (oC)

Frequency 1.0 Hz

XN-Ch2.0

XN-MgA2.0XN-Ba2.0

XN-P1.0

XN-Ca2.0

Figure 4.80 Temperature dependence of loss tangent (tan δ) of XNBR cured with 2.0x stoichiometry of various metal compounds at frequency 1.0 Hz.

Page 197: Akron 1158067153

179

CHAPTER V

CONCLUSIONS

1. The cure behavior of XNBR vulcanized by MgO depends greatly on both

concentration and specific surface area of MgO. Cure rate increases with increasing

both specific surface area and concentration.

2. Tensile properties of XNBR-MgO vulcanizates improve greatly with increased

amounts of MgO to the point where all carboxyl groups are completely neutralized,

and slightly change thereafter. The effect of surface area is not significant.

3. For XNBR-MgO compounds, neutralization occurs during mixing and continues

during storage.

4. ODR, tensile, swelling and ATR-IR results suggest that neutralization of XNBR by

MgO requires an equimolar amount of acidity and MgO. The proposed mechanisms

are 1) MgO reacts with carboxyl groups (RCOOH) to give the magnesium

hydroxycarboxylate salt, RCOOMgOH, 2) This salt reacts bimolecularly to form the

magnesium carboxylate salt, RCOOMgOOCR and Mg(OH)2.

5. Ca(OH)2 and BaO give similar effect to MgO on cure and mechanical properties of

XNBR compounds, while CaO gives similar results to thermally cured XNBR.

6. Crosslink density increases with increasing amounts of crosslinking agents, except for

the case of CaO.

Page 198: Akron 1158067153

180

7. The temperature-tan δ plot reveals an additional peak at a higher temperature in

addition to the glass-rubber transition in all ionically crosslinked systems, but not in

covalently crosslinked vulcanizates. The peak shifts to higher temperatures with

increasing concentration of curing agents.

8. The strength of ionic crosslinks increases in a predictable manner with the decrease in

size of the cations as followed: Mg++ > Ca++ > Ba++ ions.

9. The wave number of asymmetric carbonyl stretching of the carboxylate anion shifts

to lower values with increasing cation mass and size.

10. The ionic transition seems to shift to higher temperatures with increasing cation mass

and size.

Page 199: Akron 1158067153

181

REFERENCES

1. I. G. Farbenindustrie, FR 701,102, (June 21, 1933).

2. W. L. Semon, US 2,395,017, (February 19, 1946).

3. H. P. Brown, US 2,626,248, (January 20, 1953).

4. H. P. Brown, and N. G. Duke, "Carboxylic Hycar Latex – A New High Strength Elastomers", Rubber World, 130, 784, (1954).

5. H. P. Brown, and C. F. Gibbs, "Carboxylic Elastomers", Industrial and Engineering Chemistry, 47, 1006, (1955).

6. H. P. Brown, "Carboxylic Elastomers", Rubber Chemistry and Technology, 30, 1347, (1957).

7. C. F. Lufter, and N. G. Duke, "Properties of Hycar 1072 – A Carboxylic Nitrile Rubber", Industrial and Engineering Chemistry; Chemical and Engineering Data Series, 3, 297, (1958).

8. H. P. Brown, "Crosslinking Reactions of Carboxylic Elastomers", Rubber Chemistry and Technology, 36, 931, (1963).

9. C. L. Bryant, "Carboxylated Nitrile Rubbers", Journal of the Institution of the Rubber Industry, 4, 202, (1970).

10. D. K. Jenkins, and E. W. Duck, 'Carboxylated Elastomers' in "Ionic Polymers", L. Holliday, Ed., Applied Science Publisher, London, 1975, p. 173.

11. R. J. Weir, "Carboxylated Nitrile Rubber – A High Performance Elastomer", Rubber India, 32, 13, (1980).

12. W. J. MacKnight, and R. D. Lunberg, "Elastomeric Ionomers", Rubber Chemistry and Technology, 57, 652, (1984).

13. J. R. Dunn, 'Carboxylated Rubber' in "Handbook of Elastomers", 2nd ed., A. K. Bowmick, and H. L. Stephens, Eds., Marcel Dekker, New York, 2001, p. 561.

Page 200: Akron 1158067153

182

14. C. E. Frank, G. Kraus, and A. J. Haefner, "Butadiene-Methacrylic Acid Copolymers as Rubber-to-Steel Adhesives", Industrial and Engineering Chemistry, 44, 1600, (1952).

15. B. F. Goodrich, GB 723,456 (February 9, 1955).

16. J. Green, and E. F. Sverdrup, "Carboxylic Rubbers from Scrap Vulcanized Rubber", Industrial and Engineering Chemistry, 48, 2138, (1956).

17. T. F. Mika, "The Use of Epoxy Resins in Synthetic Rubber Compositions", Journal of Applied Chemistry, 6, 365, (1956).

18. S. K. Chakraborty, and S. K. De, "Epoxy Resin-Cured Carboxylated Nitrile Rubber", Journal of Applied Polymer Science, 27, 4561, (1982).

19. I. Mladenov, I. A. Tutorskii, and B. A. Dodgakin, "Reactions of Carboxylated Butadiene-Styrene Rubbers and Their Mixtures with ε-Caprolactam Induced by γ-Radiation", Journal of Polymer Science, 61, 235, (1962).

20. B. A. Dolgoplosk, E. I. Tinyakova, V. N. Reikh, T. G. Zhuravleva, and G. P. Belonovskaya, "Carboxylic Rubbers. I. Synthesis and Structures ", Rubber Chemistry and Technology, 32, 321, (1959).

21. S. K. Chakraborty, "Effect of the Curing System on the Technical Properties and the Network Structure of Carboxylated Nitrile Rubbers", Kautschuk Gummi Kunststoffe, 36, 461, (1983).

22. G. F. Beekman, and M. Hastbacka, "Altering CSM, XNBR Compounds with Magnesia", Rubber and Plastics News, December 1, 1986, p. 14.

23. K.-C. Hua, "Effect of Zinc Oxide Particle Size on the Mechanical Properties of the Caboxylated Nitrile and Carboxylated Styrene-Butadiene Rubbers", Ph.D. Dissertation, The University of Akron, 2003.

24. G. R. Hamed, and K.-C. Hua, "Effect of ZnO Particle Size on the Curing of Carboxylated NBR and Carboxylated SBR", Rubber Chemistry and Technology, 77, 214, (2004).

25. P. H. Starmer, "Effect of Metal Oxides on the Properties of Carboxyl Nitrile Rubber Vulcanisates", Plastics and Rubber Processing and Applications, 9, 209, (1988).

26. G. P. Blackmon, R. Krista, and J. Dewar, "The Effect of Vulcanization Systems on Bin Storage Stability, Physical and Cure Properties of Carboxylated Nitrile Rubbers", The Fall Meeting of the Rubber Division, ACS, Cleveland, Ohio, 2003.

27. J. C. Baker, "Improved Processing of Carboxylated Elastomers Using Zinc Peroxide", The Fall Meeting of the Rubber Division, ACS, 1966.

Page 201: Akron 1158067153

183

28. V. L. Hallenbeck, "Scorch Control of Carboxylic Elastomers", Rubber Chemistry and Technology, 46, 78, (1973).

29. J. C. Baker, US 3,403,136, (September 24, 1968).

30. N. D. Zakharov, and T. A. Shadricheva, "Influenec of Acids on Scorching and Vulcanization of Carboxylic Rubbers", Rubber Chemistry and Technology, 36, 575, (1963).

31. E. H. Andrews, "Reinforcement of Rubber by Fillers", Rubber Chemistry and Technology, 36, 325, (1963).

32. G. Kraus, "Reinforcement of Elastomers by Carbon Black", Die Angewandte Makromolekulare Chemie, 60/61, 215, (1977).

33. G. Holden, E. T. Bishop, and N. R. Legge, "Thermoplastic Elastomers", Proceeding of the International Rubber Conference, Brighton, 1967.

34. N. Nagata, T. Sato, T. Fujii, and Y. Saito, "Structure and Mechanical Properties of Hydrogenated NBR/Zinc Dimethacrylate Vulcanizates", Journal of Applied polymer Science: Applied Polymer Symposium, 53, 103, (1994).

35. C. W. Lantman, W. J. MacKnight, and R. D. Lundberg, "Structural Properties of Ionomers", Annual Review of Material science, 19, 295, (1989).

36. G. Kraus, 'Reinforcement of Elastomers by Particulate Fillers' in "Science and Technology of Rubber", F. R. Eirich, Ed., Academic Press, New York, 1978, p. 339.

37. G. Hamed, and S. Hatfield, "On the Role of Bound Rubber in Carbon Black Reinforcement", Rubber Chemistry and Technology, 62, 143, (1989).

38. G. Hamed, "Reinforcement of Rubber", Rubber Chemistry and Technology, 73, 524, (2000).

39. N. R. Legge, "Thermoplastic Elastomers", Rubber Chemistry and Technology, 60, G83, (1987).

40. G. Hamed, and B. H. Park, "The Mechanism of Carbon Black Reinforcement of SBR and NR Vulcanizates", Rubber Chemistry and Technology, 72, 946, (1999).

41. A. Nomura, J. Takano, A. Toyoda, and Y. Saito, "Structural Analysis of Very High Strength HNBR/(Zinc dimethacrylate) Composites", The 142nd Fall Meeting of the Rubber Division, ACS, Nashville, Tennesse, November 3-6, 1992.

42. H. W. Greensmith, L. Mullins, and A. G. Thomas, 'Strength of Rubbers' in "The Chemistry and Physics of Rubber-Like Substances", L. Bateman, Ed., Maclaren and Sons, London, 1963, p. 249.

Page 202: Akron 1158067153

184

43. A. N. Gent, 'Strength of Elastomers' in "Science and Technology of Rubber", F. R. Eirich, Ed., Academic Press, New Yourk, 1978, p. 419.

44. F. Bueche, "The Tensile Strength of Elastomers According to Current Theories", Rubber Chemistry and Technology, 32, 1269, (1959).

45. A. M. Bueche, "The Ultimate Properties of Simple Elastomers", Journal of Polymer Science, 19, 275, (1956).

46. G. R. Taylor, and S. R. Darin, "The Tensile Strength of Elastomers", Journal of Polymer Science, 17, 511, (1955).

47. F. Bueche, "Tensile Strength of Rubbers", Journal of Polymer Science, 24, 189, (1957).

48. P. J. Flory, "Principles of Polymer Chemistry", Cornell University Press, Ithaca, 1953, p. 485.

49. L. M. Epstein, and R. P. Smith, "The Tensile Strength of GR-S Elastomers", Transactions of the Society of Rheology, 2, 219, (1958).

50. T. L. Smith, and W. H. Chu, "Ultimate Tensile Properties of Elastomers. VII. Effect of Crosslink Density on TimeTemperature Dependence", Journal of Polymer Science: Part A-2, 10, 133, (1972).

51. L. Bateman, J. I. Cunneen, C. G. Moore, L. Mullins, and A. G. Thomas, 'Vulcanizate Structure and Physical and Aging Properties' in "The Chemistry and Physics of Rubber-Like Substances", L. Bateman, Ed., Maclaren and Sons, London, 1963, p. 715.

52. A. V. Tobolsky, and P. F. Lyons, "Tensile Strength of Rubbers", Journal of Polymer Science: Part A-2, 6, 1561, (1968).

53. A. V. Tobolsky, P. F. Lyons, and N. Hata, "Ionic Clusters in High-Strength Carboxylic Rubbers", Macromolecules, 1, 515, (1968).

54. R. Longworth, and D. J. Vaughan, "Physical Structure of Ionomers", Nature, 218, 85, (1968).

55. S. Bonotto, and E. F. Bonner, "Effect of Ion Valency on the Bulk Physical Properties of Salts of Ethylene-Acrylic Acid Copolymers", Macromolecules, 1, 510, (1968).

56. A. Eisenberg, "Clustering of Ions in Organic Polymers. A Theoretical Approach", Macromolecules, 3, 147, (1970).

57. F. C. Wilson, R. Longworth, and D. J. Vaughan, "The Physical Structure of Ionomers III. X-Ray Diffraction Studies", Polymer Preprint, 9, 505, (1968).

Page 203: Akron 1158067153

185

58. R. B. Moore, D. Bittencourt, M. Gauthier, C. E. Williams, and A. Eisenberg, "Small Angle X-Ray Scattering Investigations of Ionomers with Variable-Length Side Chains", Macromolecules, 24, 1376, (1991).

59. K. Sato, "Ionic Crosslinking of Carboxylated SBR", Rubber Chemistry and Technology, 56, 942, (1983).

60. C. L. Marx, J. A. Koutsky, and S. L. Cooper, "Morphology of Butadiene Methacrylic Acid Copolymers", Journal of Polymer Science: Polymer Letters, 9, 167, (1971).

61. J. H. Laurer, and K. I. Winey, "Direct Imaging of Ionic Aggregates in Zn-Neutralized Poly(ethylene-co-methacrylic acid) Copolymers", Macromolecules, 31, 9106, (1998).

62. U. K. Mandal, D. K. Tripathy, and S. K. De, "Moving Die Rheometry and Dynamic Mechanical Studies on the Effect of Reinforcing Carbon Black Filler on Ionomer Formation During Crosslinking of Carboxylated Nitrile Rubber by Zinc Oxide", Polymer, 34, 3832, (1993).

63. J. S. Kim, G. Wu, and A. Eisenberg, "Viscoelastic Properties of Poly(styrene-co-acrylate) and Poly(vinylcyclohexane-co-acrylate) Ionomers", Macromolecules, 27, 814, (1994).

64. K. Sato, and G. C. Blackshaw, "Dynamic Properties of Carboxylated Nitrile Rubber", Proceedings of the International Rubber Conference, Tokyo, Japan, 1985.

65. P. C. Painter, B. A. Brozoski, and M. M. Coleman, "FTIR Studies of Calcium and Sodium Ionomers Derived from an Ethylene-Methacrylic Acid", Journal of Polymer Science: Polymer Physics Edition, 20, 1069, (1982).

66. B. A. Brozoski, M. M. Coleman, and P. C. Painter, "Local Structures in Ionomer Multiplets. A Vibrational Spectroscopic Analysis", Macromolecules, 17, 230, (1984).

67. C. L. Marx, D. F. Caulfield, and S. L. Cooper, "Morphology of Ionomers", Macromolecules, 6, 344, (1973).

68. D. J. Yarusso, and S. L. Cooper, "Microstructure of Ionomers: Interpretation of Small-Angle X-ray Scattering Data", Macromolecules, 16, 1871, (1983).

69. W. J. MacKnight, W. P. Taggart, and R. S. Stein, "A Model for the Structure of Ionomers", Journal of Polymer Science: Symposium, 45, 113, (1974).

70. A. Eisenberg, and J. S. Kim, "Introduction to Ionomers", Wiley-Interscience, New York, 1998, p. 43.

71. A. Eisenberg, B. Hird, and R. B. Moore, "A New Multiplet-Cluster Model for the Morphology of Random Ionomers", Macromolecules, 23, 4098, (1990).

Page 204: Akron 1158067153

186

72. W. Cooper, "Metallic Salts of Copolymers of Butadiene and Unsaturated Acids: Some New Types of "Bouncing Putty"", Chemistry and Industry, 24, 1741, (1955).

73. W. Cooper, "Copolymers of Butadiene and Unsaturated Acids: Crosslinking by Metal Oxides", Journal of Polymer Science, 28, 195, (1958).

74. W. Cooper, "Ionic Interchange Reactions with Metal Salts of Copolymers from Butadiene and Unsaturated Acids", Journal of Polymer Science, 28, 628, (1958).

75. J. C. Halpin, and F. Bueche, "Fracture Properties of Carboxylic Rubbers", Journal of Polymer Science: Part A, 3, 3935, (1965).

76. A. K. Bhowmick, and S. K. De, "Effect of Curing Temperature on the Technical Properties of Nitrile Rubber and Carboxylated Nitrile Rubber", Rubber Chemistry and Technology, 53, 107, (1980).

77. S. K. Chakraborty, A. K. Bhowmick, and S. K. De, "Structure-Property Relations of Carboxylated Nitrile Rubber", Journal of Applied Polymer Science, 26, 4011, (1981).

78. E. P. Otocka, and F. R. Eirich, "Ionic Bonding in Butadiene Copolymers. I. Network Formation", Journal of Polymer Science: Part A-2, 6, 921, (1968).

79. L. Ibarra, and M. Alzorriz, "Ionic Elastomers Based on Carboxylated Nitrile Rubber (XNBR) and Zinc Peroxide: Influence of Carboxylic Group Content on Properties", Journal of Applied Polymer Science, 84, 605, (2002).

80. L. Ibarra, and M. Alzorriz, "Vulcanization of carboxylated nitrile rubber (XNBR) by zinc peroxide", Polymer International, 49, 580, (1999).

81. N. D. Zakharov, "Vulcanization of Carboxylic Rubbers", Rubber Chemistry and Technology, 36, 568, (1963).

82. B. A. Dolgoplosk, V. N. Reikh, E. I. Tinyakova, A. E. Kalaus, Z. A. Koryushenko, and E. G. Sladkevich, "Carboxylic Rubbers. II. Properties of Cured Stocks", Rubber Chemistry and Technology, 32, 328, (1959).

83. L. Ibarra, and M. Alzorriz, "Ionic Elastomers Based on Carboxylated Nitrile Rubber and Calcium Oxide", Journal of Applied Polymer Science, 87, 805, (2003).

84. M. R. Tant, L. N. Venkateshwaran, J. H. Song, S. Subramanian, G. L. Wilkes, P. Charlier, and R. Jerome, "Structure and Properties of Carboxylato-Telechelic Polyisoprene", Polymer, 33, 1347, (1992).

85. P. Antony, and S. K. De, "Ionic Thermoplastics Elastomers: A Review", Journal of Macromolecular Science: Polymer review, C41, 41, (2001).

Page 205: Akron 1158067153

187

86. R. J. Weir, and R. C. Burkey, "Compounding Carboxylated Nitrile Rubber", Rubber and Plastics News, February 16, 1981, p. 30.

87. F. G. Shaheen, and D. C. Grimm, "Carboxylated Nitrile Rubber", The Spring Meeting of the Rubbber Division, ACS, Los Angeles, CA, April, 24, 1985.

88. S. K. Chakraborty, and S. K. De, "Silica- and Clay-Reinforced Carboxylated Nitrile Rubber Vulcanized by a Mixed Crosslinking System", Rubber Chemistry and Technology, 55, 990, (1982).

89. U. K. Mandal, and D. K. Tripathy, "Dynamic Mechanical Properties of an Ionic Elastomer Based on Carboxylated Nitrile Rubber", Kautschuk Gummi Kunststoffe, 50, 630, (1997).

90. U. K. Mandal, D. K. Tripathy, and S. K. De, "Effect of Silica Filler on Dynamic Mechanical Properties of Ionic Elastomer Based on Carboxylated Nitrile Rubber", Journal of Applied Polymer Science, 55, 1185, (1995).

91. J. S. Kim, and A. Eisenberg, 'Ion Aggregation and Its Effect on Ionomer Properties' in "Ionomers: Characterization, Theory, and Applications", S. Schlick, Ed., CRC Press, London, 1996, p. 7.

92. H. S. Makowski, and R. D. Lundberg, 'Plasticization of Metal Sulfonate-Containing EPDM with Stearic Acid Derivatives' in "Ions in Polymers", A. Eisenberg, Ed., ACS, Washington D.C., 1980, p. 37.

93. U. K. Mandal, D. K. Tripathy, and S. K. De, "Dynamic Mechanical Spectroscopic Studies on Plasticization of an Ionic Elastomer Based on Carboxylated Nitrile Rubber by Ammonia", Polymer, 37, 5739, (1996).

94. U. K. Mandal, D. K. Tripathy, and S. K. De, "Effect of Zinc Stearate on the Rheology of Ionic Elastomer Based on Carboxylated Nitrile Rubber", Polymer Engineering and Science, 36, 283, (1996).

95. ASTM D638, "Standard Test Method for Tensile Properties of Plastics".

96. ASTM D2084, "Standard Test Method for Rubber Property – Vulcanization Using Oscillating Disk Cure Meter".

97. M. Mooney, "A Theory of Large Elastic Deformation", Journal of Applied Physics, 11, 582, (1940).

98. R. S. Rivlin, "Large Elastic Deformations of Isotropic Materials. I. Fundamental Concepts", Philosophical Transactions of the Royal Society of London: Series A, Mathematical and Physical Sciences, 240, 459, (1948).

Page 206: Akron 1158067153

188

99. R. S. Rivlin, and D. W. Saunders, "Large Elastic Deformations of Isotropic Materials. I. Experiments on the Deformation of Rubbers", Philosophical Transactions of the Royal Society of London: Series A, Mathematical and Physical Sciences, 243, 251, (1951).

100. P. J. Flory, and J. J. Rhener, "Statistical Mechanics of Cross-Linked Polymer Networks. II. Swelling ", Journal of Chemical Physics, 11, 521, (1943).

101. C. J. Sheehan, and A. L. Bisio, "Polymer/Solvent Interaction Parameters", Rubber Chemistry and Technology, 39, 149, (1966).

102. G. R. Hamed, and T. Tulyapitak, "Curing of Carboxylated Nitrile Rubber by MgO", The Spring Technical Meeting of the Rubber Division, ACS, San Antonio, Texas, May 16-18, 2005.

103. J. B. Class, "A Review of the Fundamentals of Crosslinking of Peroxides", Rubber World, 220, 35, (1999).

104. P. R. Dluzneski, "Poxide Vulcanization of Elastomers", Rubber Chemistry and Technology, 74, 451, (2001).

105. J. Y. Lee, P. C. Painter, and M. M. Coleman, "Hydrogen Bonding in Polymer Blends. 3. Blends Involving Polymers Containing Methacrylic Acid and Ether Groups", Macromolecules, 21, 346, (1988).

106. D. H. Grant, and N. Grassie, "The Thermal Decomposition of Polymethacrylic Acid", Polymer, 1, 125, (1960).

107. A. F. Holleman, and E. Wiberg, "Inorganic Chemistry", 34th ed., Academic Press, San Diego, 2001, p. 1756.

108. S. Bagrodia, and G. L. Wilkes, "Comments on the Effect of Cation Type on Ionomer Properties", Polymer Bulletin, 12, 389, (1984).

109. G. R. Hamed, 'Materials and Compounds' in "Engineering with Rubber: How to Design Rubber Components", 2nd ed., A. N. Gent, Ed., Hanser, New York, 2001, p. 11.

110. D. O. Hummel, "Atlas of Polymer and Plastics Analysis: Vol. 2 Plastics, Fibres, Rubbers, Resins; Starting and Auxiliary Materials, Degradation Products", 2nd ed., VCH Publishers, New York, 1988.

111. U. K. Mandal, "Ionic Elastomer Based on Carboxylated Nitrile Rubber: Infrared Spectral Analysis", Polymer International, 49, 1653, (2000).

112. A. Fiorenza, and G. Bonomi, "Identification of Elastomers by Infrared Spectrophotometry", Rubber Chemistry and Technology, 36, 1129, (1963).

Page 207: Akron 1158067153

189

113. J. F. O'Keefe, "Identification of Polymers by IR Spectroscopy", Rubber World, 230, 27, (2004).

114. A. Szilagyi, V. Izvekov, and I. Vancso-Szmercsanyi, "Coordination Polymers. VIII. IR Spectroscopic Studies on the Mechanism of the Reaction between Polyesters and Zinc Oxide", Journal of Polymer Science: Polymer Chemistry Edition, 18, 2803, (1980).

115. K. Han, and H. L. Williams, "lonomers: Two Formation Mechanisms and Models", Journal of Applied Polymer Science, 42, 1845, (1991).

116. U. K. Mandal, D. K. Tripathy, and S. K. De, "Effect of Carbon Black Fillers on Dynamic Mechanical Properties of Ionic Elastomer Based on Carboxylated Nitrile Rubber", Plastics, Rubber, and Composites Processing and Applications, 24, 19, (1995).

117. L. Ibarra, and M. Alzorriz, "Mechanical and Dynamic Properties of Ionic Elastomers Obtained from Carboxylated Nitrile Rubber and Zinc Peroxide", Kautschuk Gummi Kunststoffe, 53, 415, (2000).

118. K. Sakamoto, W. J. MacKnight, and R. S. Porter, "Dynamic and Steady-Shear Melt Rheology of an Ethylene-Methacrylic Acid Copolymer and Its Salts", Journal of Polymer Science: Part A-2, 8, 277, (1970).

119. B. Hird, and A. Eisenberg, "Sizes and Stabilities of Multiplets and Clusters in Carboxylated and Sulfonated Styrene Ionomers", Macromolecules, 25, 6466, (1992).

120. P. Vanhoorne, and R. A. Register, "Low-Shear Melt Rheology of Partially-Neutralized Ethylene-Methacrylic Acid Ionomers", Macromolecules, 29, 598, (1996).

121. J. S. Kim, 'Ionomers' in "Encyclopedia of Polymer Science and Technology, Vol. 6", 3rd ed., H. F. Mark, Ed., John Wiley & Sons, New Jersy, 2003, p. p600.

Page 208: Akron 1158067153

190

APPENDICES

Page 209: Akron 1158067153

191

APPENDIX A

CURE PROPERTIES

Table A.1 Cure properties of XNBR cured with different magnesium oxides at 165 oC

Property XNBR XN-MgA0.5

XN-MgA1.0

XN-MgA1.5

XN-MgA2.0

XN-MgA3.0

XN-MgA4.0

XN-MgA5.0

ML (dN.m) MH (dN.m) tS2 (min)

7.3 14.5 6.0

6.6 13.2 28.6

7.6 28.6 4.0

8.4 47.2 57.1

8.5 57.1 3.5

9.9 61.2 2.5

10.2 64.9 2.5

11.6 67.1 2.0

Property XNBR XN-MgB0.5

XN-MgB1.0

XN-MgB1.5

XN-MgB2.0

XN-MgB3.0

XN-MgB4.0

XN-MgB5.0

ML (dN.m) MH (dN.m) tS2 (min)

7.3 14.5 6.0

6.6 13.0 50.0

6.9 24.9 22.0

7.0 38.9 14.0

7.0 47.0 11.0

7.7 56.1 8.0

8.8 62.3 3.5

10.5 66.9 2.5

Property XNBR XN-MgC0.5

XN-MgC1.0

XN-MgC1.5

XN-MgC2.0

XN-MgC3.0

XN-MgC4.0

XN-MgC5.0

ML (dN.m) MH (dN.m) tS2 (min)

7.3 14.5 6.0

6.6 10.7

120.0

6.5 19.9 50.0

6.5 33.9 25.0

6.7 42.2 18.0

6.6 53.2 10.0

7.9 61.4

5

8.6 65.3

3

Page 210: Akron 1158067153

192

Table A.2 Cure properties of XNBR cured by dicumyl peroxide at 165 oC

Property XNBR XN-P0.25

XN-P0.5

XN-P0.75

XN-P1.0

XN-P1.5

XN-P2.0

XN-P3.0

ML (dN.m) MH (dN.m) MH - ML (dN.m) tS2 (min)

7.3 14.5 3.5 6.0

9.1 19.0 9.9 6.0

8.8 31.1 22.3 3.0

8.4 41.3 32.9 2.5

8.5 52.6 44.1 2.0

8.0 71.8 63.8 1.8

8.8 88.0 79.2 1.5

9.1 98.8 89.7 1.3

Table A.3 Cure properties of XNBR cured with calcium oxide at 165 oC

Property XNBR XN-Ca0.5

XN-Ca1.0

XN-Ca1.5

XN-Ca2.0

XN-Ca3.0

XN-Ca4.0

XN-Ca5.0

ML (dN.m) MH (dN.m) tS2 (min)

10.8 38.2

6

9.1 32.6 65

9.6 36.0 45

8.5 36.1 55

10.5 37.4 40

9.2 36.0 55

9.1 38.5 55

8.8 39.8 53

Table A.4 Cure properties of XNBR cured with calcium hydroxide at 165 oC

Property XNBR XN-Ch0.5

XN-Ch1.0

XN-Ch1.5

XN-Ch2.0

XN-Ch3.0

XN-Ch4.0

XN-Ch5.0

ML (dN.m) MH (dN.m) tS2 (min)

10.8 24.5

6

10.1 18.7 14

8.9 29.6 13

10.0 74.0 10

10.2 76.5 10.5

8.7 87.8 6.5

9.5 91.9 2.5

11.0 94.9 3.0

Table A.5 Cure properties of XNBR cured with barium oxide at 165 oC

Property XNBR XN-Ba0.5

XN-Ba1.0

XN-Ba1.5

XN-Ba2.0

XN-Ba3.0

XN-Ba4.0

XN-Ba5.0

ML (dN.m) MH (dN.m) tS2 (min)

10.8 24.5

6

8.8 18.8

9

9.9 27.4 15

10.9 34.7 20

9.9 44.8

7

11.0 54.3

8

9.6 65.1

5

11.6 75.2

6

Page 211: Akron 1158067153

193

APPENDIX B

TENSILE PROPERTIES

Table B.1 Tensile properties at room temperature (~25 oC) of thermally cured XNBR

Cure time Properties

60 min 120 min 240 min 500 min 1000 min

25 % Mod. (MPa) 0.28 ± 0.01 0.34 ± 0.01 0.31 ± 0.01 0.34 ± 0.01 0.35 ± 0.01

50 % Mod. (MPa) 0.37 ± 0.01 0.46 ± 0.01 0.47 ± 0.01 0.50 ± 0.01 0.52 ± 0.01

75 % Mod. (MPa) 0.42 ± 0.01 0.51 ± 0.01 0.55 ± 0.01 0.59 ± 0.01 0.62 ± 0.01

100 % Mod. (MPa) 0.44 ± 0.01 0.54 ± 0.01 0.61 ± 0.01 0.64 ± 0.01 0.69 ± 0.01

150 % Mod. (MPa) 0.46 ± 0.01 0.56 ± 0.01 0.66 ± 0.01 0.72 ± 0.01 0.78 ± 0.01

200 % Mod. (MPa) 0.48 ± 0.01 0.58 ± 0.01 0.71 ± 0.02 0.79 ± 0.01 0.86 ± 0.01

250 % Mod. (MPa) 0.49 ± 0.01 0.59 ± 0.01 0.76 ± 0.02 0.85 ± 0.01 0.94 ± 0.01

300 % Mod. (MPa) 0.50 ± 0.01 0.61 ± 0.01 0.81 ± 0.03 0.92 ± 0.02 1.03 ± 0.02

350 % Mod.(MPa) 0.51 ± 0.01 0.63 ± 0.01 0.87 ± 0.03 1.01 ± 0.03 1.14 ± 0.03

400 % Mod. (MPa) 0.52 ± 0.01 0.65 ± 0.01 0.94 ± 0.03 1.10 ± 0.04 1.27 ± 0.04

450 % Mod. (MPa) 0.54 ± 0.02 0.68 ± 0.01 1.02 ± 0.04 1.22 ± 0.05 1.42 ± 0.06

500 % Mod. (MPa) 0.55 ± 0.02 0.71 ± 0.01 1.11 ± 0.05 1.36 ± 0.06 1.63 ± 0.08

TS (MPa) > 4.20 ± 0.56 6.46 ± 0.15 7.79 ± 0.13 8.19 ± 0.18 9.08 ± 0.14

EB (%) > 1655 ± 52 1420 ± 14 939 ± 13 827 ± 3 751 ± 2

Page 212: Akron 1158067153

194

Table B.2 Tensile properties at room temperature (~25 oC) of XN-MgA vulcanizates (cure time 30 min)

Properties XNBR XN-

MgA0.5 XN-

MgA1.0 XN-

MgA1.5 XN-

MgA2.0 XN-

MgA3.0 XN-

MgA4.0 XN-

MgA5.0

25 % Mod. (MPa)

0.29 ± 0.01

0.83 ± 0.03

0.86 ± 0.02

1.40 ± 0.12

1.40 ± 0.03

1.43 ± 0.06

1.52 ± 0.05

1.61 ± 0.03

50 % Mod. (MPa)

0.41 ± 0.01

1.10 ± 0.04

1.24 ± 0.02

2.08 ± 0.16

2.20 ± 0.04

2.28 ± 0.05

2.41 ± 0.05

2.52 ± 0.04

75 % Mod. (MPa)

0.46 ± 0.01

1.25 ± 0.05

1.48 ± 0.02

2.55 ± 0.19

2.80 ± 0.04

2.94 ± 0.05

3.09 ± 0.05

3.29 ± 0.06

100 % Mod. (MPa)

0.48 ± 0.01

1.34 ± 0.05

1.66 ± 0.02

2.95 ± 0.22

3.33 ± 0.05

3.56 ± 0.07

3.72 ± 0.06

3.97 ± 0.08

150 % Mod. (MPa)

0.50 ± 0.01

1.43 ± 0.05

1.98 ± 0.02

3.55 ± 0.30

4.38 ± 0.09

4.83 ± 0.12

5.04 ± 0.10

5.41 ± 0.14

200 % Mod. (MPa)

0.50 ± 0.02

1.50 ± 0.05

2.31 ± 0.02

4.60 ± 0.30

5.60 ± 0.12

6.32 ± 0.15

6.53 ± 0.16

7.01 ± 0.20

250 % Mod. (MPa)

0.50 ± 0.02

1.57 ± 0.04

2.72 ± 0.04

5.77 ± 0.42

7.19 ± 0.17

8.10 ± 0.26

8.35 ± 0.21

8.94 ± 0.28

300 % Mod. (MPa)

0.50 ± 0.02

1.64 ± 0.04

3.24 ± 0.07

7.34 ± 0.59

9.44 ± 0.28

10.4 ± 0.4

10.8 ± 0.3

11.5 ± 0.4

350 % Mod. (MPa)

0.50 ± 0.02

1.73 ± 0.05

3.94 ± 0.12

9.67 ± 0.86

12.7 ± 0.4

13.8 ± 0.6

13.6 ± 1.5

14.9 ± 0.7

400 % Mod. (MPa)

0.50 ± 0.02

1.86 ± 0.04

4.93 ± 0.19

13.0 ± 1.3

17.6 ± 0.6

18.3 ± 1.0

19.0 ± 0.7

19.4 ± 1.1

450 % Mod. (MPa)

0.51 ± 0.03

2.03 ± 0.05

6.53 ± 0.36

17.8 ± 1.9

24.1 ± 0.8

24.6 ± 1.5

25.1 ± 1.0

25.3 ± 1.3

500 % Mod. (MPa)

0.52 ± 0.03

2.27 ± 0.06

9.13 ± 0.55

24.2 ± 2.6

32.5 ± 1.2

32.5 ± 2.1

32.8 ± 1.0

32.7 ± 1.8

TS (MPa)

Not break >3.60

13.5 ± 0.5

28.5 ± 0.9

43.5 ± 2.4

46.5 ± 1.7

48.4 ± 0.5

48.6 ± 2.4

52.8 ± 2.0

EB (%)

Not break >1600

830 ± 12

674 ± 13

614 ± 15

572 ± 11

584 ± 12

583 ± 15

611 ± 14

Page 213: Akron 1158067153

195

Table B.3 Tensile properties at room temperature (~25 oC) of XN-MgA vulcanizates (cure time 120 min)

Properties XNBR XN-MgA0.5

XN-MgA1.0

XN-MgA1.5

XN-MgA2.0

XN-MgA3.0

XN-MgA4.0

XN-MgA5.0

25 % Mod. (MPa)

0.34 ± 0.01

0.57 ± 0.01

0.95 ± 0.01

1.25 ± 0.03

1.43 ± 0.01

1.72 ± 0.05

1.79 ± 0.01

1.72 ± 0.02

50 % Mod. (MPa)

0.46 ± 0.01

0.76 ± 0.01

1.40 ± 0.03

1.85 ± 0.03

2.07 ± 0.01

2.42 ± 0.07

2.51 ± 0.01

2.59 ± 0.01

75 % Mod. (MPa)

0.51 ± 0.01

0.86 ± 0.02

1.70 ± 0.03

2.20 ± 0.03

2.68 ± 0.01

3.13 ± 0.09

3.25 ± 0.03

3.22 ± 0.03

100 % Mod. (MPa)

0.54 ± 0.01

0.93 ± 0.02

1.90 ± 0.03

2.51 ± 0.05

3.14 ± 0.02

3.66 ± 0.11

3.79 ± 0.03

3.95 ± 0.04

150 % Mod. (MPa)

0.56 ± 0.01

1.02 ± 0.03

2.39 ± 0.04

3.28 ± 0.07

4.14 ± 0.02

4.98 ± 0.16

5.17 ± 0.05

5.27 ± 0.08

200 % Mod. (MPa)

0.58 ± 0.01

1.08 ± 0.03

2.88 ± 0.06

4.21 ± 0.12

5.46 ± 0.05

6.67 ± 0.22

6.79 ± 0.08

6.89 ± 0.12

250 % Mod. (MPa)

0.59 ± 0.01

1.15 ± 0.03

3.48 ± 0.07

5.51 ± 0.20

7.21 ± 0.05

8.74 ± 0.27

8.94 ± 0.15

8.97 ± 0.35

300 % Mod. (MPa)

0.61 ± 0.01

1.23 ± 0.03

4.30 ± 0.09

7.30 ± 0.31

9.67 ± 0.10

11.6 ± 0.3

11.7 ± 0.3

11.7 ± 0.3

350 % Mod. (MPa)

0.63 ± 0.01

1.33 ± 0.04

5.51 ± 0.13

10.4 ± 0.5

13.0 ± 0.2

15.9 ± 0.5

15.6 ± 0.3

15.8 ± 0.5

400 % Mod. (MPa)

0.65 ± 0.01

1.46 ± 0.04

7.30 ± 0.16

15.6 ± 0.9

18.1 ± 0.3

21.8 ± 0.6

21.2 ± 0.5

21.1 ± 0.7

450 % Mod. (MPa)

0.68 ± 0.01

1.63 ± 0.06

10.3 ± 0.3

23.1 ± 1.5

25.6 ± 0.3

30.2 ± 1.0

29.4 ± 0.8

29.2 ± 0.8

500 % Mod. (MPa)

0.71 ± 0.01

1.85 ± 0.06

15.0 ± 0.4

33.2 ± 2.2

35.3 ± 0.3

40.0 ± 1.4

39.1 ± 1.0

38.0 ± 1.0

TS (MPa)

6.46 ± 0.15

15.9 ± 0.5

34.8 ± 1.2

44.7 ± 1.0

48.3 ± 1.1

51.5 ± 1.7

51.4 ± 1.2

51.5 ± 1.0

EB (%)

1420 ± 14

872 ± 7

635 ± 5

549 ± 3

553 ± 16

548 ± 8

552 ± 7

561 ± 20

Page 214: Akron 1158067153

196

Table B.4 Tensile properties at room temperature (~25 oC) of XN-MgB vulcanizates (cure time 30 min)

Properties XNBR XN-

MgB0.5 XN-

MgB1.0 XN-

MgB1.5 XN-

MgB2.0 XN-

MgB3.0 XN-

MgB4.0 XN-

MgB5.0

25 % Mod. (MPa)

0.29 ± 0.01

0.44 ± 0.02

0.70 ± 0.02

0.94 ± 0.01

1.15 ± 0.02

1.49 ± 0.04

1.63 ± 0.02

1.58 ± 0.05

50 % Mod. (MPa)

0.41 ± 0.01

0.62 ± 0.02

1.01 ± 0.03

1.35 ± 0.01

1.71 ± 0.05

2.35 ± 0.03

2.50 ± 0.06

2.57 ± 0.03

75 % Mod. (MPa)

0.46 ± 0.01

0.70 ± 0.02

1.19 ± 0.01

1.62 ± 0.03

2.12 ± 0.06

3.00 ± 0.05

3.18 ± 0.06

3.36 ± 0.02

100 % Mod. (MPa)

0.48 ± 0.01

0.76 ± 0.01

1.32 ± 0.01

1.83 ± 0.04

2.47 ± 0.0

3.58 ± 0.07

3.83 ± 0.07

4.09 ± 0.03

150 % Mod. (MPa)

0.50 ± 0.01

0.83 ± 0.01

1.53 ± 0.01

2.22 ± 0.05

3.15 ± 0.09

4.72 ± 0.09

5.04 ± 0.07

5.49 ± 0.05

200 % Mod. (MPa)

0.50 ± 0.02

0.90 ± 0.02

1.78 ± 0.05

2.65 ± 0.08

3.91 ± 0.12

5.96 ± 0.13

6.40 ± 0.10

6.99 ± 0.05

250 % Mod. (MPa)

0.50 ± 0.02

0.96 ± 0.02

2.01 ± 0.02

3.17 ± 0.11

4.84 ± 0.16

7.53 ± 0.18

8.02 ± 0.18

8.73 ± 0.05

300 % Mod. (MPa)

0.50 ± 0.02

1.03 ± 0.02

2.35 ± 0.03

3.86 ± 0.13

6.09 ± 0.26

9.63 ± 0.29

10.2 ± 0.3

11.0 ± 0.1

350 % Mod. (MPa)

0.50 ± 0.02

1.13 ± 0.03

2.82 ± 0.05

4.79 ± 0.21

7.90 ± 0.39

12.6 ± 0.4

13.3 ± 0.4

14.2 ± 0.2

400 % Mod. (MPa)

0.50 ± 0.02

1.26 ± 0.04

3.47 ± 0.08

6.17 ± 0.32

10.7 ± 0.6

17.0 ± 0.8

17.8 ± 0.7

18.8 ± 0.3

450 % Mod. (MPa)

0.51 ± 0.03

1.45 ± 0.05

4.49 ± 0.15

8.36 ± 0.52

14.7 ± 0.9

22.9 ± 1.5

23.8 ± 1.1

25.1 ± 0.4

500 % Mod. (MPa)

0.52 ± 0.03

1.70 ± 0.07

6.21 ± 0.24

11.7 ± 0.8

20.4 ± 1.3

30.6 ± 1.6

31.6 ± 1.4

33.2 ± 0.5

TS (MPa)

Not break >3.60

5.96 ± 0.24

17.2 ± 0.5

26.0 ± 1.3

33.3 ± 1.5

40.1 ± 0.5

45.8 ± 1.6

46.0 ± 2.1

EB (%)

Not break >1600

724 ± 12

642 ± 12

619 ± 11

586 ± 11

553 ± 11

575 ± 15

567 ± 14

Page 215: Akron 1158067153

197

Table B.5 Tensile properties at room temperature (~25 oC) of XN-MgB vulcanizates (cure time 120 min)

Properties XNBR XN-MgB0.5

XN-MgB1.0

XN-MgB1.5

XN-MgB2.0

XN-MgB3.0

XN-MgB4.0

XN-MgB5.0

25 % Mod. (MPa)

0.34 ± 0.01

0.62 ± 0.02

0.97 ± 0.03

1.24 ± 0.04

1.41 ± 0.03

1.61 ± 0.05

1.71 ± 0.09

1.71 ± 0.07

50 % Mod. (MPa)

0.46 ± 0.01

0.87 ± 0.02

1.42 ± 0.01

1.88 ± 0.05

2.20 ± 0.06

2.62 ± 0.12

2.68 ± 0.12

2.87 ± 0.08

75 % Mod. (MPa)

0.51 ± 0.01

1.01 ± 0.01

1.71 ± 0.03

2.35 ± 0.06

2.80 ± 0.07

3.49 ± 0.17

3.48 ± 0.12

3.74 ± 0.08

100 % Mod. (MPa)

0.54 ± 0.01

1.10 ± 0.01

1.95 ± 0.02

2.76 ± 0.07

3.32 ± 0.07

4.28 ± 0.22

4.20 ± 0.12

4.60 ± 0.08

150 % Mod. (MPa)

0.56 ± 0.01

1.25 ± 0.02

2.39 ± 0.03

3.54 ± 0.10

4.33 ± 0.08

5.76 ± 0.32

5.61 ± 0.16

6.19 ± 0.10

200 % Mod. (MPa)

0.58 ± 0.01

1.38 ± 0.02

2.87 ± 0.03

4.42 ± 0.16

5.48 ± 0.11

7.37 ± 0.46

7.13 ± 0.22

7.85 ± 0.12

250 % Mod. (MPa)

0.59 ± 0.01

1.54 ± 0.03

3.47 ± 0.04

5.53 ± 0.27

6.93 ± 0.14

9.32 ± 0.68

9.01 ± 0.34

9.82 ± 0.17

300 % Mod. (MPa)

0.61 ± 0.01

1.74 ± 0.05

4.27 ± 0.06

7.09 ± 0.42

8.85 ± 0.23

11.9 ± 0.9

11.5 ± 0.5

12.3 ± 0.3

350 % Mod. (MPa)

0.63 ± 0.01

2.01 ± 0.05

5.43 ± 0.07

9.42 ± 0.65

11.7 ± 0.4

15.5 ± 1.5

15.1 ± 0.7

15.8 ± 0.5

400 % Mod. (MPa)

0.65 ± 0.01

2.38 ± 0.08

7.21 ± 0.09

13.0 ± 0.9

15.8 ± 0.7

20.5 ± 2.2

20.2 ± 1.0

20.7 ± 0.8

450 % Mod. (MPa)

0.68 ± 0.01

2.92 ± 0.15

10.1 ± 0.2

18.1 ± 1.4

21.5 ± 1.1

27.1 ± 3.3

27.0 ± 1.5

27.2 ± 1.2

500 % Mod. (MPa)

0.71 ± 0.01

3.77 ± 0.24

14.3 ± 0.4

24.9 ± 2.1

28.9 ± 1.2

35.4 ± 4.4

35.7 ± 1.9

35.5 ± 1.6

TS (MPa)

6.46 ± 0.15

13.6 ± 0.5

25.2 ± 0.9

38.7 ± 2.8

42.3 ± 2.7

47.9 ± 1.7

49.3 ± 1.9

51.3 ± 1.3

EB (%)

1420 ± 14

687 ± 12

588 ± 9

579 ± 9

571 ± 14

563 ± 28

568 ± 13

579 ± 15

Page 216: Akron 1158067153

198

Table B.6 Tensile properties at room temperature (~25 oC) of XN-MgC vulcanizates (cure time 30 min)

Properties XNBR XN-

MgC0.5 XN-

MgC1.0 XN-

MgC1.5 XN-

MgC2.0 XN-

MgC3.0 XN-

MgC4.0 XN-

MgC5.0

25 % Mod. (MPa)

0.29 ± 0.01

0.32 ± 0.01

0.46 ± 0.01

0.83 ± 0.02

0.98 ± 0.02

1.34 ± 0.04

1.34 ± 0.02

1.56 ± 0.03

50 % Mod. (MPa)

0.41 ± 0.01

0.44 ± 0.01

0.62 ± 0.01

1.14 ± 0.02

1.45 ± 0.02

2.05 ± 0.03

1.91 ± 0.03

2.28 ± 0.04

75 % Mod. (MPa)

0.46 ± 0.01

0.49 ± 0.01

0.71 ± 0.01

1.34 ± 0.02

1.75 ± 0.02

2.57 ± 0.04

2.36 ± 0.03

2.85 ± 0.04

100 % Mod. (MPa)

0.48 ± 0.01

0.52 ± 0.01

0.76 ± 0.01

1.49 ± 0.02

1.99 ± 0.02

3.04 ± 0.04

2.88 ± 0.03

3.38 ± 0.03

150 % Mod. (MPa)

0.50 ± 0.01

0.53 ± 0.01

0.81 ± 0.01

1.73 ± 0.03

2.45 ± 0.03

3.99 ± 0.06

3.80 ± 0.03

4.54 ± 0.03

200 % Mod. (MPa)

0.50 ± 0.02

0.53 ± 0.01

0.85 ± 0.01

2.00 ± 0.03

2.96 ± 0.03

5.05 ± 0.09

4.85 ± 0.02

5.72 ± 0.03

250 % Mod. (MPa)

0.50 ± 0.02

0.53 ± 0.01

0.90 ± 0.02

2.31 ± 0.04

3.58 ± 0.04

6.36 ± 0.14

6.19 ± 0.02

7.43 ± 0.04

300 % Mod. (MPa)

0.50 ± 0.02

0.53 ± 0.01

0.95 ± 0.02

2.71 ± 0.05

4.39 ± 0.07

8.07 ± 0.22

7.99 ± 0.03

9.55 ± 0.09

350 % Mod. (MPa)

0.50 ± 0.02

0.53 ± 0.01

1.02 ± 0.02

3.24 ± 0.07

5.49 ± 0.13

10.5 ± 0.4

10.8 ± 0.1

12.6 ± 0.1

400 % Mod. (MPa)

0.50 ± 0.02

0.54 ± 0.01

1.12 ± 0.03

3.96 ± 0.10

7.12 ± 0.26

14.0 ± 0.7

15.0 ± 0.1

17.3 ± 0.2

450 % Mod. (MPa)

0.51 ± 0.03

0.54 ± 0.01

1.24 ± 0.04

5.04 ± 0.15

9.64 ± 0.44

19.1 ± 1.1

21.6 ± 0.2

23.7 ± 0.2

500 % Mod. (MPa)

0.52 ± 0.03

0.54 ± 0.01

1.41 ± 0.05

6.77 ± 0.29

13.4 ± 0.7

25.6 ± 1.6

30.6 ± 0.5

32.2 ± 0.2

TS (MPa)

Not break >3.60

Not break >1.70

6.3 ± 0.2

18.9 ± 2.0

28.9 ± 1.4

40.4 ± 1.3

37.3 ± 1.2

46.4 ± 1.4

EB (%)

Not break >1600

Not break >1600

834 ± 14

651 ± 14

624 ± 13

588 ± 17

537 ± 7

570 ± 6

Page 217: Akron 1158067153

199

Table B.7 Tensile properties at room temperature (~25 oC) of XN-MgC vulcanizates (cure time 120 min)

Properties XNBR XN-MgC0.5

XN-MgC1.0

XN-MgC1.5

XN-MgC2.0

XN-MgC3.0

XN-MgC4.0

XN-MgC5.0

25 % Mod. (MPa)

0.34 ± 0.01

0.46 ± 0.01

0.89 ± 0.02

1.20 ± 0.02

1.33 ± 0.03

1.54 ± 0.03

1.46 ± 0.06

1.51 ± 0.03

50 % Mod. (MPa)

0.46 ± 0.01

0.63 ± 0.01

1.30 ± 0.03

1.78 ± 0.02

2.06 ± 0.03

2.43 ± 0.04

2.16 ± 0.05

2.32 ± 0.02

75 % Mod. (MPa)

0.51 ± 0.01

0.73 ± 0.01

1.56 ± 0.03

2.17 ± 0.02

2.61 ± 0.03

3.14 ± 0.07

2.72 ± 0.05

2.96 ± 0.02

100 % Mod. (MPa)

0.54 ± 0.01

0.78 ± 0.01

1.75 ± 0.03

2.51 ± 0.02

3.10 ± 0.04

3.78 ± 0.08

3.23 ± 0.05

3.60 ± 0.02

150 % Mod. (MPa)

0.56 ± 0.01

0.84 ± 0.02

2.10 ± 0.02

3.17 ± 0.03

4.08 ± 0.06

5.05 ± 0.12

4.33 ± 0.05

4.84 ± 0.03

200 % Mod. (MPa)

0.58 ± 0.01

0.89 ± 0.02

2.49 ± 0.04

3.91 ± 0.06

5.18 ± 0.09

6.42 ± 0.17

5.58 ± 0.07

6.19 ± 0.05

250 % Mod. (MPa)

0.59 ± 0.01

0.94 ± 0.02

2.97 ± 0.05

4.81 ± 0.09

6.54 ± 0.13

8.09 ± 0.26

7.06 ± 0.10

7.70 ± 0.08

300 % Mod. (MPa)

0.61 ± 0.01

1.00 ± 0.02

3.60 ± 0.07

5.98 ± 0.14

8.36 ± 0.22

10.3 ± 0.5

9.04 ± 0.13

9.79 ± 0.13

350 % Mod. (MPa)

0.63 ± 0.01

1.06 ± 0.03

4.46 ± 0.12

7.72 ± 0.21

11.0 ± 0.4

13.4 ± 0.8

11.1 ± 0.1

12.7 ± 0.2

400 % Mod. (MPa)

0.65 ± 0.01

1.16 ± 0.03

5.77 ± 0.20

10.3 ± 0.4

14.8 ± 0.6

17.6 ± 1.1

16.4 ± 0.2

16.6 ± 0.4

450 % Mod. (MPa)

0.68 ± 0.01

1.29 ± 0.04

7.88 ± 0.34

14.2 ± 0.6

20.1 ± 1.03

23.3 ± 1.6

21.1 ± 0.3

22.2 ± 0.6

500 % Mod. (MPa)

0.71 ± 0.01

1.45 ± 0.06

11.1 ± 0.5

19.7 ± 0.9

26.9 ± 1.5

30.7 ± 2.2

30.2 ± 0.3

29.2 ± 0.7

TS (MPa)

6.46 ± 0.15

7.5 ± 1.4

26.6 ± 1.2

36.2 ± 2.0

43.3 ± 2.2

49.6 ± 2.0

47.4 ± 0.2

48.5 ± 0.9

EB (%)

1420 ± 14

815 ± 14

637 ± 5

605 ± 15

592 ± 12

600 ± 13

581 ± 3

607 ± 6

Page 218: Akron 1158067153

200

Table B.8 Tensile properties at room temperature (~25 oC) of XNBR-peroxide vulcanizates (cure time 60 min)

Properties XNBR XN-

P0.25 XN-

P0.50 XN-

P0.75 XN-P1.0

XN-P1.5

XN-P2.0

XN-P3.0

25 % Mod. (MPa)

0.28 ± 0.01

0.35 ± 0.01

0.38 ± 0.01

0.37 ± 0.01

0.38 ± 0.01

0.43 ± 0.01

0.48 ± 0.02

0.57 ± 0.01

50 % Mod. (MPa)

0.37 ± 0.01

0.48 ± 0.01

0.52 ± 0.01

0.53 ± 0.01

0.58 ± 0.01

0.67 ± 0.01

0.75 ± 0.02

0.93 ± 0.02

75 % Mod. (MPa)

0.42 ± 0.01

0.55 ± 0.01

0.60 ± 0.01

0.63 ± 0.01

0.69 ± 0.01

0.82 ± 0.01

0.92 ± 0.02

1.22 ± 0.02

100 % Mod. (MPa)

0.44 ± 0.01

0.59 ± 0.01

0.66 ± 0.01

0.71 ± 0.01

0.77 ± 0.02

0.94 ± 0.01

1.07 ± 0.02

1.48 ± 0.04

150 % Mod. (MPa)

0.46 ± 0.01

0.64 ± 0.01

0.74 ± 0.01

0.82 ± 0.02

0.91 ± 0.02

1.16 ± 0.02

1.34 ± 0.01

2.04 ± 0.06

200 % Mod. (MPa)

0.48 ± 0.01

0.68 ± 0.01

0.82 ± 0.01

0.94 ± 0.02

1.04 ± 0.02

1.41 ± 0.02

1.64 ± 0.01 -

250 % Mod. (MPa)

0.49 ± 0.01

0.72 ± 0.01

0.90 ± 0.02

1.08 ± 0.04

1.20 ± 0.04

1.70 ± 0.02

2.04 ± 0.03 -

300 % Mod. (MPa)

0.50 ± 0.01

0.76 ± 0.02

1.00 ± 0.03

1.24 ± 0.05

1.42 ± 0.04

2.06 ± 0.04 - -

350 % Mod. (MPa)

0.51 ± 0.01

0.82 ± 0.02

1.11 ± 0.03

1.43 ± 0.05

1.66 ± 0.04

2.52 ± 0.06 - -

400 % Mod. (MPa)

0.52 ± 0.01

0.86 ± 0.01

1.23 ± 0.03

1.63 ± 0.05

1.96 ± 0.04

3.18 ± 0.10 - -

450 % Mod. (MPa)

0.54 ± 0.02

0.92 ± 0.02

1.36 ± 0.04

1.86 ± 0.05

2.33 ± 0.02 - - -

500 % Mod. (MPa)

0.55 ± 0.02

0.98 ± 0.02

1.51 ± 0.04

2.11 ± 0.05

2.78 ± 0.03 - - -

TS (MPa)

> 4.20 ± 0.56

8.33 ± 0.25

8.06 ± 0.22

5.66 ± 0.93

4.59 ± 0.26

3.58 ± 0.42

2.51 ± 0.09

2.35 ± 0.09

EB (%)

> 1655 ± 52

1424 ± 12

1010 ± 4

769 ± 32

611 ± 12

421 ± 20

292 ± 9

173 ± 10

Page 219: Akron 1158067153

201

Table B.9 Tensile properties at room temperature (~25 oC) of XNBR-CaO vulcanizates (cure time 1000 min)

Properties XNBR XN-Ca0.5

XN-Ca1.0

XN-Ca1.5

XN-Ca2.0

XN-Ca3.0

XN-Ca4.0

XN-Ca5.0

25 % Mod. (MPa)

0.35 ± 0.01

0.41 ± 0.02

0.38 ± 0.02

0.43 ± 0.01

0.43 ± 0.02

0.42 ± 0.01

0.44 ± 0.02

0.44 ± 0.01

50 % Mod. (MPa)

0.52 ± 0.01

0.58 ± 0.01

0.55 ± 0.01

0.59 ± 0.01

0.60 ± 0.01

0.60 ± 0.01

0.63 ± 0.02

0.65 ± 0.01

75 % Mod. (MPa)

0.62 ± 0.01

0.68 ± 0.01

0.65 ± 0.01

0.69 ± 0.01

0.70 ± 0.01

0.71 ± 0.01

0.73 ± 0.01

0.77 ± 0.01

100 % Mod. (MPa)

0.69 ± 0.01

0.75 ± 0.01

0.71 ± 0.01

0.75 ± 0.01

0.76 ± 0.01

0.78 ± 0.01

0.80 ± 0.02

0.84 ± 0.01

150 % Mod. (MPa)

0.78 ± 0.01

0.87 ± 0.01

0.81 ± 0.01

0.83 ± 0.01

0.86 ± 0.01

0.87 ± 0.01

0.90 ± 0.02

0.94 ± 0.01

200 % Mod. (MPa)

0.86 ± 0.01

0.98 ± 0.01

0.89 ± 0.01

0.91 ± 0.01

0.95 ± 0.01

0.96 ± 0.01

0.98 ± 0.02

1.04 ± 0.01

250 % Mod. (MPa)

0.94 ± 0.01

1.10 ± 0.01

0.98 ± 0.01

0.98 ± 0.01

1.03 ± 0.01

1.05 ± 0.01

1.07 ± 0.01

1.13 ± 0.01

300 % Mod. (MPa)

1.03 ± 0.02

1.22 ± 0.01

1.07 ± 0.01

1.06 ± 0.02

1.13 ± 0.01

1.14 ± 0.01

1.18 ± 0.02

1.23 ± 0.01

350 % Mod. (MPa)

1.14 ± 0.03

1.35 ± 0.01

1.18 ± 0.01

1.15 ± 0.02

1.22 ± 0.01

1.24 ± 0.01

1.28 ± 0.03

1.35 ± 0.02

400 % Mod. (MPa)

1.27 ± 0.04

1.51 ± 0.01

1.31 ± 0.01

1.26 ± 0.03

1.34 ± 0.02

1.35 ± 0.02

1.39 ± 0.01

1.47 ± 0.02

450 % Mod. (MPa)

1.42 ± 0.06

1.70 ± 0.01

1.46 ± 0.01

1.39 ± 0.04

1.47 ± 0.02

1.49 ± 0.02

1.52 ± 0.03

1.60 ± 0.03

500 % Mod. (MPa)

1.63 ± 0.08

1.98 ± 0.02

1.64 ± 0.01

1.55 ± 0.05

1.63 ± 0.02

1.66 ± 0.03

1.70 ± 0.03

1.77 ± 0.03

TS (MPa)

9.08 ± 0.14

8.21 ± 0.18

7.71 ± 0.15

7.95 ± 0.20

7.52 ± 0.60

6.96 ± 0.33

7.16 ± 0.11

7.59 ± 0.09

EB (%)

751 ± 2

692 ± 10

746 ± 3

769 ± 15

764 ± 19

772 ± 8

766 ± 7

785 ± 5

Page 220: Akron 1158067153

202

Table B.10 Tensile properties at room temperature (~25 oC) of XNBR-Ca(OH)2 vulcanizates (cure time 240 min)

Properties XNBR XN-Ch0.5

XN-Ch1.0

XN-Ch1.5

XN-Ch2.0

XN-Ch3.0

XN-Ch4.0

XN-Ch5.0

25 % Mod.

(MPa)

0.31 ± 0.01

0.53 ± 0.01

1.67 ± 0.08

2.39 ± 0.01

2.28 ± 0.02

2.37 ± 0.08

2.47 ± 0.09

2.61 ± 0.05

50 % Mod.

(MPa)

0.47 ± 0.01

0.73 ± 0.01

2.67 ± 0.06

4.14 ± 0.02

4.20 ± 0.02

4.44 ± 0.04

4.69 ± 0.04

5.05 ± 0.12

75 % Mod.

(MPa)

0.55 ± 0.01

0.86 ± 0.02

3.36 ± 0.06

5.54 ± 0.05

5.64 ± 0.01

6.13 ± 0.04

6.50 ± 0.06

6.97 ± 0.09

100 % Mod.

(MPa)

0.61 ± 0.01

0.96 ± 0.02

3.93 ± 0.08

6.68 ± 0.03

6.92 ± 0.03

7.49 ± 0.5

7.94 ± 0.04

8.46 ± 0.09

150 % Mod.

(MPa)

0.66 ± 0.01

1.12 ± 0.03

5.01 ± 0.10

8.69 ± 0.03

9.17 ± 0.11

9.76 ± 0.07

10.2 ± 0.1

10.8 ± 0.1

200 % Mod.

(MPa)

0.71 ± 0.02

1.26 ± 0.04

6.23 ± 0.12

10.8 ± 0.1

11.2 ± 0.1

11.9 ± 0.1

12.4 ± 0.1

13.0 ±0.2

250 % Mod.

(MPa)

0.76 ± 0.02

1.43 ± 0.05

7.87 ± 0.17

13.3 ± 0.1

13.8 ± 0.1

14.6 ± 0.1

15.0 ± 0.1

15.7 ± 0.2

300 % Mod.

(MPa)

0.81 ± 0.03

1.63 ± 0.05

10.3 ± 0.3

16.9 ± 0.1

17.3 ± 0.1

18.2 ± 0.1

18.5 ± 0.1

19.3 ± 0.3

350 % Mod.

(MPa)

0.87 ± 0.03

1.88 ± 0.06

14.1 ± 0.4

22.0 ± 0.1

22.2 ± 0.1

23.2 ± 0.2

23.1 ± 0.2

24.2 ±0.5

400 % Mod.

(MPa)

0.94 ± 0.03

2.20 ± 0.07

19.9 ± 0.5

29.0 ± 0.3

28.7 ± 0.1

29.7 ± 0.2

29.4 ± 0.2

30.3 ± 0.7

450 % Mod.

(MPa)

1.02 ± 0.04

2.64 ± 0.08

27.8 ± 0.7

37.6 ± 0.5

36.9 ± 0.4

37.8 ± 0.5

36.8 ± 0.3

37.8 ± 0.7

500 % Mod.

(MPa)

1.11 ± 0.05

3.31 ± 0.12

38.0 ± 0.9 - 46.3

± 0.4 46.9 ± 0.3

45.1 ± 0.5

46.1 ± 0.8

TS

(MPa)

7.79 ± 0.13

14.9 ± 0.8

41.8 ± 0.8

45.1 ± 1.7

50.9 ± 1.0

48.2 ± 0.7

47.5 ± 0.6

48.1 ± 0.2

EB

(%)

939 ± 13

712 ± 12

516 ± 3

486 ± 11

522 ± 7

507 ± 4

514 ± 2

511 ± 3

Page 221: Akron 1158067153

203

Table B.11 Tensile properties at room temperature (~25 oC) of XNBR-BaO vulcanizates (cure time 240 min)

Properties XNBR XN-Ba0.5

XN-Ba1.0

XN-Ba1.5

XN-Ba2.0

XN-Ba3.0

XN-Ba4.0

XN-Ba5.0

25 % Mod. (MPa)

0.31 ± 0.01

0.45 ± 0.01

0.64 ± 0.01

0.84 ± 0.01

0.96 ± 0.02

1.29 ± 0.04

2.03 ± 0.06

2.38 ± 0.01

50 % Mod. (MPa)

0.47 ± 0.01

0.66 ± 0.01

0.99 ± 0.01

1.32 ± 0.01

1.58 ± 0.04

2.21 ± 0.04

3.19 ± 0.06

3.76 ± 0.05

75 % Mod. (MPa)

0.55 ± 0.01

0.77 ± 0.01

1.18 ± 0.01

1.62 ± 0.03

1.99 ± 0.03

2.83 ± 0.06

4.06 ± 0.06

4.85 ± 0.07

100 % Mod. (MPa)

0.61 ± 0.01

0.84 ± 0.01

1.32 ± 0.01

1.85 ± 0.02

2.29 ± 0.05

3.34 ± 0.07

4.84 ± 0.04

5.80 ± 0.09

150 % Mod. (MPa)

0.66 ± 0.01

0.96 ± 0.02

1.57 ± 0.01

2.28 ± 0.03

2.87 ± 0.07

4.29 ± 0.09

6.34 ± 0.08

7.60 ± 0.12

200 % Mod. (MPa)

0.71 ± 0.02

1.06 ± 0.02

1.83 ± 0.01

2.73 ± 0.04

3.50 ± 0.08

5.34 ± 0.13

8.02 ± 0.10

9.58 ± 0.18

250 % Mod. (MPa)

0.76 ± 0.02

1.17 ± 0.03

2.13 ± 0.01

3.29 ± 0.07

4.28 ± 0.11

6.71 ± 0.16

10.2 ± 0.1

12.1 ± 0.2

300 % Mod. (MPa)

0.81 ± 0.03

1.30 ± 0.04

2.54 ± 0.01

4.04 ± 0.10

5.34 ± 0.13

8.58 ± 0.23

13.1 ± 0.1

15.4 ± 0.2

350 % Mod. (MPa)

0.87 ± 0.03

1.47 ± 0.05

3.08 ± 0.02

5.06 ± 0.18

6.86 ± 0.22

11.3 ± 0.3

17.1 ± 0.1

19.7 ± 0.3

400 % Mod. (MPa)

0.94 ± 0.03

1.69 ± 0.04

3.84 ± 0.03

6.58 ± 0.24

9.26 ± 0.32

15.3 ± 0.4

22.5 ± 0.4

25.3 ± 0.5

450 % Mod. (MPa)

1.02 ± 0.04

1.98 ± 0.05

5.05 ± 0.07

9.09 ± 0.37

12.9 ± 0.3

20.9 ± 0.2 - -

500 % Mod. (MPa)

1.11 ± 0.05

2.38 ± 0.06

7.10 ± 0.19

13.1 ± 0.5

18.3 ± 0.3 - - -

TS (MPa)

7.79 ± 0.13

7.52 ± 0.20

12.8 ± 0.6

19.2 ± 0.5

21.7 ± 1.6

26.6 ± 0.7

25.6 ± 0.3

26.0 ± 0.3

EB (%)

939 ± 13

684 ± 5

580 ± 9

554 ± 4

526 ± 8

491 ± 3

424 ± 5

407 ± 3

Page 222: Akron 1158067153

204

APPENDIX C

MOLECULAR TRANSITION TEMPERATURE

Table C.1 Molecular transition temperatures of XN-MgB vulcanizates at frequency 1.0 Hz

Vulcanizates Tg (oC) Ionic transition temperature range (oC)

XN-MgB1.0 -20 10 - 85

XN-MgB2.0 -19 15 - 105

XN-MgB3.0 -18 30 - 110

Table C.2 Molecular transition temperatures of XN-MgC vulcanizates at frequency 1.0 Hz

Vulcanizates Tg (oC) Ionic transition temperature range (oC)

XN-MgC1.0 -21 15 - 85

XN-MgC2.0 -17 20 - 100

XN-MgC3.0 -17 25 - 105