Anda di halaman 1dari 15

GEO Specialty Chemicals

Specialty Additives Group


340 Mathers Road
Ambler, PA 19002-3420
USA

Technical Information
Bulletin: ORC-101

DI-CUP and VUL-CUP Peroxide


Fundamentals of Crosslinking
INTRODUCTION
Sulfur has been used since the early 1800s for the vulcanization of rubber. Peroxides, on the other hand,
have only been used for this purpose since the mid-1900s. In 1914, a Russian researcher first discovered
that benzoyl peroxide could crosslink natural rubber, but it was observed that the resulting vulcanizates
had lower strength and poorer heat resistance than did sulfur vulcanizates. It was not until 1950 that it
was found that di-tert-butyl peroxide gives better properties to vulcanizates than does benzoyl peroxide.
Unfortunately, di-tert-butyl peroxide is volatile, resulting in excessive loss during incorporation into the
rubber and, therefore, poor efficiency. Soon after this observation, Hercules developed DI-CUP
dicumyl peroxide, the first commercially available peroxide to combine high efficiency and good
vulcanizate properties with low cost, thus providing a peroxide crosslinker with broad-spectrum utility.
Today, many elastomers and plastics are commercially crosslinked using organic peroxides.
Widespread acceptability of a chemical crosslinker occurs if the crosslinker satisfies a number of
conditions. According to Braden, Fletcher, and McSweeney1 these are as follows:
-

It must be safe to handle, non-irritating, and non-toxic at processing temperatures.


It must react to give crosslinks as the only modification of the polymer.
Its decomposition characteristics must be such as to give a rapid cure at
approximately 150C, with no tendency to scorch at processing temperatures.
It must be nonvolatile to prevent loss during mixing.
It must be soluble in rubber and plastics.
Neither the peroxide nor its decomposition products must accelerate aging of the
rubber or plastic.
It must be effective in the presence of reinforcing fillers.

Peroxides used for crosslinking fall into four general classes: dialkyl peroxides, diacyl peroxides, peroxy
esters, and peroxy ketals. Each is used for its particular reactivity in a particular crosslinking system.
1

Transactions of Institute of the Rubber Industry, 30, 44 (1954).

ORC-101
Page 2 of 15

The most predominantly used products are in the dialkyl class, specifically dicumyl peroxide (Di-Cup)
and a,a-bis (t-butyl peroxy) diisopropylbenzene (Vul-Cup).
The primary criteria for selecting a peroxide is the activation temperature, or the temperature at which
crosslinking will proceed at a satisfactory rate. Peroxides with lower activation temperatures, which are
indicated by lower half-life temperatures, will provide fast cure times, but will have poor scorch
resistance. High activation temperature peroxides will provide good scorch time, but have longer cure
times.
Other factors to consider when selecting the most cost-effective peroxide to use are:
Efficiency - or the fraction of peroxide active oxygen that participates in the
crosslink function
Volatility - will the peroxide volatilize during processing
Di-Cup and Vul-Cup are solids that melt at relatively low temperatures and are essentially non-volatile at
processing conditions. The chemical structures of these two peroxides provide for near 100% efficiency:
all of the active oxygen contributes to crosslinking.
The development of peroxide vulcanization systems poses questions: Why should a polymer compounder
be interested in a peroxide cure? Will a peroxide-cured vulcanizate perform in a different manner than a
sulfur-cured vulcanizate? Partial answers can be seen upon examining the differences between the two
crosslinks.
From the standpoint of thermal stability, the crosslink resulting from peroxide cure has a bond energy of
343.2 kJoules and is as stable as any of the carbon-carbon (C-C) bonds in the polymer backbone. The
crosslink resulting from a sulfur cure is composed of carbon-sulfur and sulfur-sulfur bonds. These have
lower bond energies (C-S, 276.2 kJoules; S-S, 205.1 kJoules) than C-C bonds and result in a weaker
crosslink.
Figure 1
Typical Peroxide-Induced Crosslink
Peroxide cures result in crosslinks of strong
carbon-to-carbon covalent bonds between
adjacent polymer chains.

C
C
Figure 2
Typical Sulfur-Induced Crosslink

C
Sulfur cures result in crosslinks of polyatomic sulfur bridges between adjacent
polymer chains.

SX
C

The stronger crosslinks formed by peroxide vulcanization provide superior heat aging and compression
set than is found in sulfur vulcanizates. In addition, peroxide vulcanizates exhibit little or no added color
while sulfur systems form dark metallic sulfides. On the other hand, sulfur vulcanization generally yields
better abrasion resistance and tear strength than do peroxide vulcanizates due to the greater flexibility of
the polysulfidic crosslink.

ORC-101
Page 3 of 15

FUNDAMENTALS OF PEROXIDE CROSS-LINKING


Polymers That Can Be Peroxide-Crosslinked:
NR
SBR
BR
IR
NBR
CR
EPM
EPDM

- natural rubber
- styrene-butadiene rubber
- polybutadiene rubber
- polyisoprene rubber
- nitrile rubber
- polychloroprene rubber
- ethylene-propylene copolymer
- ethylene-propylene terpolymer

MQ
ACM
T
AU, EU
ABS
PE
CSM
CM

- silicone rubber
- acrylic rubber
- polysulfide rubber
- polyurethane
- acrylonitrile-butadiene-styrene
- polyethylene (LDPE, EVA, etc.)
- chlorosulfonated polyethylene
- chlorinated polyethylene (CPE)

Polymers That Cannot Be Peroxide-Crosslinked:


IIR
CO, ECO
PP
PVC

- butyl rubber
- polyepichlorohydrin rubber
- polypropylene
- polyvinyl chloride

Basic Reactions:
Peroxides thermally cleave to produce two oxy radicals.
Acyl peroxides yield acyloxy radicals; alkyl peroxide

1.

ROOR 2 RO

2.

RO + H-P ROH + P

yields alkoxy radicals.


Oxy radicals are very reactive and abstract hydrogen
atoms from polymer chains, where P = polymer.
Two polymer radicals then combine to form a crosslink.
Certain polymers having pendant vinyl or other very
active groups may crosslink via a polymerization-type
crosslink (polybutadiene and vinyl silicone rubbers).

3.
4.

P + P P-P
RO+ n

[ C=C ] RO-[C-

C]n
Reaction 2 indicates the most desired reaction of oxy radicals, which will subsequently lead to
crosslinking. The oxy radical is very reactive and will abstract hydrogens from any available source.
Hydrogen abstraction from other than polymer molecules will reduce crosslinking efficiency.
The oxy radical will abstract any available hydrogen from
a compound.

5.

RO + H-R1 ROH + R1

ORC-101
Page 4 of 15

The relative ease with which hydrogens are abstracted by oxy radicals varies greatly. The following list
shows the relative abstractability of hydrogens in different functional groups:

Chemical
Group

Chemical
Structure

Phenyl

-H

(Most Difficult to Abstract)

Acid
Alcohol

Methane

Vinylic

R-O-H

105

RO

H
|
H-C-H
|
H
C=C-H

Tertiary
Carbon

R
|
R-C-H
|
R

Allylic

C=C-CH
-C-H

-N-H
-O-H

105

104
100

||

R-CO

Secondary
Carbon

(Easiest to Abstract)

106

R
|
R-C-H
|
H

Phenol

R-C-O-H

Primary
Carbon

Aniline

Radical
Structure

112

O
||

R
|
H-C-H
|
H

Benzylic

Bond Dissociation
Energy (kcal/mol)

H
|

HC |
H

C=C
R
|

HC |
H

97

R
|
R-C
|
H

91

R
|
R-C
|
R

C=C-C

88
88

-C

88

-N

87

- O

ORC-101
Page 5 of 15

REACTIONS DETRIMENTAL TO BASIC POLYMER CROSSLINKING


The following reactions are detrimental to the basic polymer crosslinking step:
Acid Cleavage
Peroxides are unstable in the presence of acids and cleave
ionically. The resulting products do not contain radicals
and, therefore, do not lead to crosslinking. This reaction
reduces peroxide crosslinking efficiency.

6.

Non-Radical
Decomposition
ROOR + Acid
Products

Curing in the Presence of Oxygen


7.
A radical that has been transferred to a polymer chain is
susceptible to oxygenation. A hydroperoxide forms readily
and then thermally decomposes in a reaction leading to
polymer degradation.

+ O2

-OOH

beta Cleavage
Certain polymers degrade in the presence of decomposing
peroxides. This is caused by beta cleavage associated with
special structural features of the polymer, notably pendant
methyl groups. Polymers that are prone to beta cleavage are
polypropylene, butyl rubber, and polyisobutylene.

8.

FUNDAMENTALS OF RATE OF PEROXIDE CURING REACTION


Since the chemistries of peroxide and sulfur vulcanization are very different, peroxide cure systems do
not respond to the technology of sulfur-accelerator vulcanizing systems. Knowledge of sulfur cures does
not lend itself to the design of a peroxide-cured stock. For this reason, a basic discussion of important
aspects of the peroxide crosslinking process is included.
Peroxide Half-Life
The main determinant of the cure rate in a peroxide-based system is the rate at which the peroxide
decomposes at a specific temperature. Of the three reaction steps involved to make a crosslink (Reactions
1, 2, and 3 mentioned previously) the thermally-induced cleavage of the peroxide bond is the slowest,
rate-determining step.
"Half-Life" is a more useful term than "rate constant" in describing peroxide crosslinking reactions. The
half-life is the time it takes one-half of any quantity of peroxide present to thermally decompose. The
time is independent of the quantity of peroxide present and is primarily dependent on temperature. In the

=O

ORC-101
Page 6 of 15

first half-life, 50% of the peroxide decomposes. During the second half-life, 50% of the remaining
peroxide decomposes. This is 25% of the original quantity, and the total amount decomposed is now
75%. The process continues and theoretically never reaches 100% decomposed. Practically, the level is
so low after seven half-lives that the effect of additional cure time is insignificant. Table I shows the
relationship between number of half-lives and the portion of the peroxide that has been decomposed.
Table I
Relationship Between Half-Life and Decomposition
Half-Lives

Amount of Original Peroxide Decomposed


or Fraction of Ultimate Cure State, %

1
2
3
4
5
6
7
8
9
10

50
75
87.5
93.75
96.9
98.4
99.2
99.6
99.8
99.9

Time/Temperature as a Function of the Peroxide Half-Life


The peroxide thermal decomposition reaction (equation 1) follows first-order kinetics. The kinetic order
was determined by partially curing compounds for varying cure times and then analyzing unreacted or
unused peroxide. The rate of peroxide thermal decomposition thus obtained coincided with the
decomposition rate as determined on an oscillating disc rheometer (ODR).

Torque

100

80

80

60

60

40

40

20
0

Residual Peroxide

4
6
Half-Lives

20

ODR Torque (lb-in)

The peroxide decomposition curve and the ODR


curve show the direct dependence of
crosslinking on peroxide decomposition.

Peroxide Remaining (%)

Figure 3
The Relationship Between Crosslinking and Peroxide Decomposition

10

Figure 3 clearly shows that a peroxide cure reaches a plateau when all of the peroxide has
decomposed. Changing the peroxide level in the compound will change the state of cure, but will not
change the time necessary to reach a plateau. For a given polymer system, cure time can be altered only
by a change in cure temperature.

ORC-101
Page 7 of 15

Cure Time as a Function of the Medium


The major factor affecting the rate of peroxide decomposition, and therefore the rate of cure, is
temperature. In practice it has been found that the medium (i.e., polymer) within which the peroxide
decomposes has a pronounced effect on the rate of peroxide decomposition. For example, the half-life of
Di-Cup R dicumyl peroxide at 160C is 3.5 min. in solution, but slows to 10 min. in PE and ethylenevinyl acetate (EVA). The half-life curves describing crosslinking rates for Vul-Cup peroxide in
polyethylene, EP rubbers, BR, NBR, SBR, NR, and IR rubbers, and in solution, are shown in Figure 4.
Half-life curves for Di-Cup are shown in Figure 5. Other commercially available peroxides have been
found to behave similarly.
Figures 4 and 5 show that the polymer affects the rate of peroxide crosslinking. The precise reasons for
this variation in rate are not well known; however, they are undoubtedly related to the viscosity and
solvent characteristics of the polymer.

FUNDAMENTALS OF CURE AND COMPOUND DESIGN


Figure 3 shows that a cure plateau is reached in any peroxide system after a cure time of seven to 10 halflives. Does this mean that all peroxide stocks should be cured this long? The answer is no. Peroxide
stocks can be cured for less time, provided the following conditions are fulfilled:
-

Excess (unreacted) peroxide is economically acceptable.

Optimum compression set resistance is not required.

Ultimate tensile strength is not required.

Post-curing can be done with an understanding of and compensation for its effect on the physical
properties of the vulcanizate.
100
60

EPM, EPDM

100

Figure 4
Half-Life of Vul-Cup
vs Temperature

60

40

40
SBR, NR

20

PE, EVA

20

10

Half-Life (minutes)

Half-Life (minutes)

Figure 5
Half-Life of Di-Cup
vs Temperature

6
0.1
2
1
0.6

EPM, EPDM

PE, EVA

10
6
0.1
2
1
0.6

Solution

0.4

0.4
Solution

0.2

0.2
NBR

0.1
Temperature
C

NBR

0.1
F

300
148

320
160

340
171

360
182

380
193

400 Temperature
C
204

300
148

320
160

340
171

360
182

380
193

400
204

ORC-101
Page 8 of 15

Figure 6
Typical Response to Cure Time for Peroxide Cure.

Compression Set, %

Tensile Strength,
psi (mPa)
2,500
(17.2)

Tensile Strength
300% Modulus

2,000
(13.8)

300% Modulus
psi (mPa)
1,500
(10.3
60
)
40

1,500
(10.3)

Compression Set

t95

20

1,000
(6.9)

500
(3.4
)

t99

Cure Time vs. Tensile Strength and Compression Set Resistance


Figure 6, on the previous page, shows typical 300% modulus, tensile strength, and
compression set resistance response to cure time for a peroxide cure. Tensile strength rapidly
attains a satisfactory value. Modulus rises idly at first and slowly comes to a plateau when
all of the peroxide has decomposed. Compression set follows the same pattern as modulus.
There appears to be very little gain in properties from t95 (time to 95% cure) to the cure
plateau. The major benefit of curing to the cure plateau is a minor gain in compression set
resistance. If optimum compression set resistance is not required, the cure for many
polymers can be terminated between t90 and t95.
Post-curing in a Circulating-Air Oven
During post-curing, peroxide decomposition can lead to either crosslinking or chain scission.
Figure 7 shows the effect of post-cure on stocks susceptible to chain scission in the presence of
oxygen, that have been press cured for increasing lengths of time. Stocks cured for short times
contain unused peroxide. Post-curing in the presence of oxygen leads to chain scission and reduced
compound performance. When the stocks are cured to a cure plateau, the peroxide is gone, and postcure causes no change in performance.

ORC-101
Page 9 of 15

Figure 7
Effect of Post-cure on Stocks Susceptible to Chain Scission

Tensile Strength

Key:
Property after Press Cure.

2,000
(13.8)

Property after Press Cure


Plus Post-Cure in Air.

1,500
(10.3)

1,000
(6.9)

Compression Set

60

40

20

Compression Set (%)

Tensile Strength, psi (MPa)

2,500
(17.2)

t95
0
Press-Cure Time (minutes)

The response of physical properties to the post-curing of partially press-cured vulcanizates varies
from polymer to polymer. The example used in Figure 7 was an EPDM stock. This elastomer
exhibits behavior common for polymers that are largely saturated.
More unsaturated polymers, such as NBR, deviate from this smooth response. Figure 8 shows the
result of post-curing NBR vulcanizates after they have been partially press cured. Note that
depending on the press-cure time, post-curing may lead to either an increase or decrease in properties
depending on the elastomer involved.

Tensile Strength

3,000
(21.0)

Key:
Property after Press Cure.

2,500
(17.2)

80

Property after Press Cure


Plus Post-Cure in Air.

60
2,000
(13.8)

1,500
(10.3)

Compression Set

40

t95
20
Press-Cure Time (minutes)

Compression Set (%)

Tensile Strength, psi (MPa)

Figure 8
Effect of Postcuring NBR After Partial Press Cure

ORC-101
Page 10 of 15

Compression Set Resistance


Compression set is a measure of a compound's inability to recover its original dimensions after being
compressed under heated conditions. Unreacted curative will contribute to poor compression set
resistance because it continues to vulcanize while the sample is being compressed. These new
crosslinks then prevent the stock from returning to its original shape.
If peroxide formulations are cured to the cure plateau, seven to 10 half-lives, no further curing is
possible and compression set resistance is at a maximum for that particular compound. Compression
set resistance in stocks cured in this manner varies directly with state of cure (or the quantity of
peroxide in the formulation). Figure 9 shows the typical response of an elastomer fully cured with
various levels of peroxide. The isolated point in the figure shows the result of under-curing a stock
and utilizing only 50% of the peroxide added.
Figure 9
Typical Response of Fully Cured Elastomer

Compound Containing
1.2 phr of Vul-Cup Peroxide
Cured to One Half-Life

50

Compression Set (%)

40

30

20

10

0
0.5

Vul-Cup (phr)

1.0

ORC-101
Page 11 of 15

Effect of Compounding Ingredients


The basic peroxide chemistry discussed earlier can be readily translated into compounding do's and
don'ts.
1. Fillers
The only significant reaction between a peroxide and a filler occurs when acidic fillers ionically
cleave the peroxide. This depletes the peroxide in an unproductive manner so that the efficiency of
the peroxide is reduced. Neutral and basic fillers have no effect on peroxides.
2. Oils and Plasticizers
All oils and plasticizers reduce the efficiency of peroxides because some of the radicals generated are
transferred to the oil and do not lead to crosslinking. The degree to which these materials reduce cure
efficiency is dependent on the level of addition and the type of oil or plasticizer used. Aromatic oils
have the greatest detrimental effect, naphthenic oils are intermediate, and paraffinic oils have the least
effect. Plasticizers such as dioctyl phthalate are similar in effect to paraffinic oils.
3. Antioxidants
All antioxidants reduce the efficiency of peroxides during vulcanization. Amine-type antioxidants are
preferred because they have the least effect on cure. Phenolic-type materials inhibit cure to a greater
extent.
Recommended Antioxidants
Agerite Resin D
Irganox 1010, 1035
Antioxidant 425, 2246

R.T. Vanderbilt Co.,Inc.


CIBA-GEIGY
Corporation
CYTEC Industries, Inc.

Santonox R
Flectol H
Polygard

FLEXSYS America LT.


FLEXSYS America LT.
Harwick Standard Distribution Co.
Uniroyal Chemical, Inc.

4. Antiozonants
These materials should be avoided in peroxide-cured formulations, if possible, because they greatly
reduce the efficiency of the peroxide.
Fundamentals for Developing Peroxide
Vulcanizate Modulus Equivalent to That of a Sulfur Vulcanizate
When developing a new formulation based on a peroxide cure, one should keep the formulation
simple. If an existing sulfur formulation is to be used, remove all ingredients that are part of the
curative package. The following basic formulation is suggested as a guide in compound
development:
Formulation
Polymer
Filler
Zinc oxide
Antioxidant
Peroxide

Parts by Weight
100
Variable
5.0
0.5
(See Below)
Di-Cup
EPDM,EPM,CM
2.4-3.2
VMQ, BR
0.5-1.0
All other polymers
1.0-2.0

Vul-Cup
1.5-2.0
0.3-0.6
0.6-1.2

ORC-101
Page 12 of 15

The two methods most often used to develop a peroxide-cured formulation having unaged modulus
values equivalent to those of a sulfur vulcanizate follow.
Method 1. Plot modulus vs. Peroxide level. The exact peroxide level corresponding to the desired
modulus can be read from the resulting curve. Four to five levels of a peroxide in a given
formulation, each sample cured seven to 10 half-lives, result in a 200% modulus response curve of
the type shown in Figure 10.

For example, a sulfur vulcanizate having a


200% modulus value of 1,250 psi would be
duplicated in a peroxide system with 0.52 phr
of the peroxide in question.

200% Modulus, psi (MPa)

Figure 10
Modulus vs. Peroxide Level

2000
(13.8)
1500
(10.3)
1000
(6.9)
500
(3.4)
0.5
1.0
Peroxide (phr)

Method 2. ODR (see next section). This method is applicable only to similar stocks and where roomtemperature physicals of both the sulfur and the peroxide stocks relate equally to their elevatedtemperature physicals.
OSCILLATING DISC RHEOMETER (ODR)
The ODR is useful in determining cure times and in developing peroxide vulcanizates with cures
equivalent to those of other systems.
Basic Rules on the Use of the ODR
The ODR should be set at an oscillation rate of 3-100 cpm and an arc of 30. A larger arc or a higher
oscillation rate will cause heat buildup in the sample. When this happens, the instrument measures
the curing profile at a higher temperature than that set, and the results are meaningless. We have
found that 900 cpm can cause a temperature differential as great as 15F (8C). For the most accurate
results, 3 cpm should be used.
Determining Crosslinking Half-Life
ODR torque is directly related to the number of crosslinks formed during peroxide vulcanization.
The crosslinking half-life can be determined from the ODR plot in two ways.
A simplified method is based on the concept that the time to 95% cure (t95) is 4.35 half-lives. The
measurement of t95 is not precise, since various compounds heat up at different rates and the macro
die takes longer than the micro die. In addition, a variable preheat period may be set on the ODR. To
make this method work, total heat-up time (including preheat time) must be subtracted from t95.

ORC-101
Page 13 of 15

GEO Specialty Chemicals experience indicates the heat-up time for the micro die is 30 seconds, and
for the macro die is 4 minutes. Thus, the 95% cure time minus the heat-up time is 4.35 half lives.
A more accurate method to determine crosslinking half-life is to plot log delta torque (torque infinitytorque at time t) vs. time t. The slope of this straight line (m) can be used to calculate half-life. This
method is independent of sample heat-up time.
Figure 11
Crosslinking Half-Life ODR Torque vs. Time

T2
ODR Torque

T1

t1

Time

t2

t = 0.301
m

Log Delta Torque

Figure 12
Crosslinking Half-Life
Log Delta Torque vs. Time

Slope = m

Time

ORC-101
Page 14 of 15

Determining the Peroxide Level for Equivalent Cures


The ODR is a convenient instrument for determining the quantity of peroxide necessary to cure a
compound to the same state of cure as an existing cure system does. This method is applicable only
when two formulations of the same stock, differing only in cure system, are compared at the same
temperature. The method is based on the fact that two compounds of the same basic recipe have
equivalent torques at the same state of cure. The procedure follows.
1. Run the cure profile of the reference stock.
2. Determine the optimum cure time from the ODR plot (time B, Figure 13).
3. Run the cure profile on the peroxide compound. The peroxide stock curve must cross the
reference stock curve. If the peroxide stock curve does not cross, additional peroxide
must be added to the formulation. The exact peroxide level added to the stock must be
known.
4. Determine the time, D, at which the peroxide stock has the same torque, C, as the
reference stock, A, cured to time B.
Figure 13
Determining the Peroxide Level Needed to Obtain an Equivalent Cure

Peroxide Stock

ODR Torque

Reference Stock
A

Time
Determine the number of crosslinking half-lives (n) by dividing time D by the half-life.
D
n = --------

t
Determine the percent peroxide reacted at time D from the number of half-lives, using
Figure 14; % reacted = 100% - % remaining.

ORC-101
Page 15 of 15

Figure 14
Log of Percent Peroxide Unreacted vs. Half-Life

Peroxide Remaining (%)

100

10

3
4
Half-Lives (n)

Calculate the quantity of peroxide used to time D by multiplying the percent reacted by the initial
level of peroxide.
P D = % x PO
This level of peroxide, P, will give an equivalent cure when cured to a cure plateau (seven to 10
half-lives).
Product Safety
Read and understand the Material Safety Data Sheet (MSDS) before using this product.
For additional information, and to place an order or sample request, call (888) 519-3883.
PRD 7/03

Anda mungkin juga menyukai