Anda di halaman 1dari 14

A Chemical Blowing Agent System

(CBAS) Based on Azodicarbonamide


G. L. A. SIMS* AND H. A. S. JAAFAR
Manchester Materials Science Centre
University of Manchester and UMIST
Grosvenor Street, Manchester M1 7HS, U.K.

INTRODUCTION

hemical blowing agents ( CBAs) are often used in the manufacture


C of cross-linked polyethylene
CBA azodicarbonamide
and PVC foams [1]. The most popular
is (ADC) which is selected for its high gas yield
and the ability to tailor the decomposition temperature by the use of
suitable activators [2]. Jaafar and Sims [3] investigated the thermal
decomposition behaviour of ADC in the absence of polymer matrix us-
ing gas rate apparatus (Figure 1) and DSC. It was found that the
decomposition behaviour was a function of particle size, heating rate,
activator type and concentration, and degree of dispersion of blowing
agent (BA) and activator. However, the rate of gas yield of activated
ADC is relatively rapid. It is surmised that more controllable foaming
may be achieved by a more gradual release of gas. Recent research in-
vestigated the so-called double blowing systems with the following
principle of foaming action. In a double mixture the &dquo;auxiliary&dquo; blow-
ing agent (usually CBA) serves as a nucleation agent and is the first to
operate prior to the main blowing agent [4].
This paper reports on the investigation of the formulation of a CBAS
in an attempt to control the rate of gas evolution and the resulting
foam structure. Commonly available CBAs (Table 1) were viewed in the
context of decomposition attributes to determine co-components which

*Author to whom all correspondence should be addressed.

JOURNAL OF CELWLAR PLASTICS Volume 30- March 1994 175


0021-955X/94/02 0175-14 $06.00/0
@1994 Technomic Publishing Co., Inc.
176 G L A. SIMS AND H. A. S. JAAFAR

Figure 1. Gas evolution rate apparatus [3].

could be blended to modify gas yield rate. Ideal properties of a co-


blowing agent would be decomposition temperature below but
reasonably close to the decomposition of ADC and good gas yield. It
may be seen from Table 1 that the most favourable BA against these pa-
rameters is 4,4’ -oxybisbenzenesulfonylhydrazide (OBSH). It is one of
the best blowing agents of the sulfonylhydrazide class [5]. Since it is a
difunctional sulfonylhydrazide it imparts no odor to plastics or rubbers
because its decomposition gives nonvolatile and nontoxic oligomers.

Table 1. Common blowing agents used for production of polyolefin foams [5].
A CBAS Based on Azodicarbonamide 177

EXPERIMENTAL PROCEDURE

Raw Materials

The ADC grade used was Genitron AC 1 from Schering Polymer Addi-
tives with a declared average particle size of 20 Am [3] (subsequently
referred to simply as AC1). OBSH was selected to be the co-blowing
agent for three reasons: it has a high heat of combustion, a high gas
yield, and its decomposition temperature is lower than ADC but nearer
than other alternatives. OBSH is a white crystalline compound with a
decomposition temperature 130-140 ° C as quoted in Table 1. However,
these data do not specify heating rate which has a strong influence. The
majority of this work was carried out at 6 ° C/min, and at this heating
rate OBSH decomposition temperature was found to be 165°C with a
total gas yield of 290 mls/g. It has the following chemical structure:

All materials used were from the same batch to avoid batch to batch
variation.

CBAS Preparation

All mixtures were blended on a weight basis. The additives were


mixed on a small scale using a high sensitivity balance capable of
reading to ±0.0001 g. One gram of AC1 was placed in each flask and
calculated as a hundred parts. Other additives were weighed into each
flask according to the specific system. Each ingredient was calculated
on the basis of parts per hundred of AC1 (phACl) or parts per hundred
of OBSH (phOBSH). Unless otherwise specified, all formulations were
investigated in the absence of polymer matrix.
A mechanical flask shaker with four arms was used to agitate the
mixed samples with identical frequency and time of agitation. In all
cases, a measured quantity of approximately 0.010 g of blowing agent,
thoroughly mixed activated blowing agent or blowing agent system,
was weighed into the decomposition tube of the gas rate apparatus

(Figure 1).
Instrumentation

A DSC Du Pont Instruments 2000 equipped with a 910 cell base was
178 G. L. A. SIMS AND H. A. S. JAAFAR

used to measure both the temperature and heat changes associated


with transitions in the materials, using an open pan, compared with an
inactive reference.
Gas evolution rate apparatus, depicted in Figure 1, was utilized to de-
termine the rate of gas evolution and total gas yield from a specific
weight of blowing agent or blowing agent mixture (approximately
0.010 g) placed in the decomposition tube. As the temperature of the
heating block increased, gas was produced and the volume recorded by
readings of the mercury scale. The silicone oil temperature was similar
to the decomposition temperature obtained by DSC; therefore the oil
temperature was used as reference. Blowing agent formulations were
decomposed in the gas apparatus at a heating rate of approximately
6 ° C/min. Mercury scale readings were converted to volume. The system
was initially calibrated by noting air expansion as a function of tem-

perature. The volume of gas liberated from the blowing agent at partic-
ular temperatures was then determined by subtracting the appropriate
air expansion calibration volume from the total gas volume. This ad-
justed figure was used to calculate the volume of gas liberated per unit
weight of BA (mls/g).
A Scanning Electron Microscope was used to investigate the cell
structure of foams expanded by activated AC1 and a preferred CBAS.

Foam Preparation

Samples of foam were prepared using a two stage compression mould-


ing technique [6]. The foam formulation consisted of 100 parts of
polymer, 15 phr total blowing agent and 1 phr dicumylperoxide (DCP)
cross-linking agent. All studies were based on EVA copolymer (UL
00206 manufactured by Exxon) with a reported melt flow index of 2
decigrams/min and 6% vinyl acetate content.
The polymer was placed into the nip of a two roll-mill where both
rolls were controlled at 115 ° C. This temperature plus the shear exerted
from the two roll-mill was sufficient to form a band. Polymer granules
were added continuously to the mill in small amounts and the nip gap

adjusted to give a rolling bank of approximately 10 mm. Subsequently


AC1 or CBAS was added in a similar way cutting and folding continu-
ously to ensure adequate mixing. DCP was added last to reduce the
risk of premature cross-linking. After sheeting off, 90 g of the formula-
tion was placed into a preheated mould (which equated to approx-
imately 103% of the volume of the cavity to ensure total filling). The
mould was placed in a compression press and subjected to 13,800 KPa
for 30 min at 160°C, during which time cross-linking and complete
A CBAS Based on Azodicarbonamide 179

decomposition of the blowing agent system was achieved. The mould


was then cooled under this pressure until the temperature fell to 30 ° C.
Following pressure release, a partially expanded blank was produced
containing blowing agent gasses in partial solution. This was imme-
diately transferred to a fan assisted oven at 135 ° C for 30 min to com-
plete the expansion.
RESULTS AND DISCUSSION

When examining the properties of AC1 and OBSH blowing agents


separately, using the gas rate apparatus, differences were seen in
decomposition behaviour as shown in Figure 2.
Blending the two blowing agents in equal proportions (AC1
100/OBSH 100) resulted in rapid gas liberation (Figure 3). It appeared
that the heat generated by the exothermic decomposition from the high
concentration of OBSH was sufficient enough to activate the ADC at a
much lower temperature ( 165-170 ° C ).
Consequently, it was inferred that to obtain gas release over a
suitable temperature range the OBSH concentration in the blowing
agent mixture must be reduced (to reduce the OBSH activation of
ADC) and the decomposition temperature of AC1 should be reduced to
within the suitable decomposition range. This was achieved by the ad-
dition of a small portion of activator that specifically activates AC 1 but
not OBSH. DSC showed that at 6 ° C/min the OBSH decomposition tem-

Figure 2. Decomposition behaviour of AC1 and OBSH (sample size = 0.010 g).
180 G. L. A. SIMS AND H. A. S. JAAFAR

Figure 3. Activation of ADC gas evolution with OBSH (formulation in parts by weight,
sample size = 0.010 g).

perature not affected by the addition of zinc oxide (ZnO) at 2 and


was
4 phOBSH; whereas it is well known that ZnO is used as a &dquo;kicker&dquo; for
ADC [2]. In an attempt to achieve more gradual gas evolution rates,
OBSH was therefore used at lower levels of addition with the inclusion
of ZnO. Decomposition performance of two formulations is shown in
Figure 4.
It may beclearly seen that both curves, whilst not optimized, showed
more gradual gas release than AC1 or OBSH alone or AC1 100/OBSH
100. Three phAC 1 ZnO appeared to activate ADC too soon so that the
OBSH exotherm virtually took ADC with it. It is notable that this com-
position is more controlled than AC1 100/OBSH 100 (Figure 3). The for-
mulation containing 2 parts of ZnO showed the more gradual decompo-
sition behaviour. These results were supported by DSC (Figure 5) that
showed OBSH and AC1 peak exotherms at 159 and 201 ° C which corre-
spond approximately to the maximum rate of yield as depicted in
Figure 4.
As the AC 1 100/OBSH 50/ZnO 2 formulation appeared to match more
closely the original aim, this system was investigated further. Previous
work [3] showed that the decomposition behaviour of ADC was strictly
dependent on heating rate. The initial results reported in this paper
were carried out at 6 ° C/min but typical production processing methods
can have considerably higher heating rates. The formulation was
therefore subjected to DSC measurements at varying heating rates.
A CBAS Based on Azodicarbonamide 181

Figure 4. The influence of ZnO on CBAS decomposition (formulations in parts by


weight, sample size = 0.010 g).

Figure 5. DSC curve for mixture AC1 100/OBSH 50/ZnO 2 at 6°C/min.


182 G. L. A. SIMS AND H. A. S. JAAFAR

Figure 6 clearly illustrated that at 30°C/min the OBSH and AC 1


decomposition exotherms had virtually merged.
The greater dependence of decomposition temperature behaviour on
varying heating rates is shown in Figure 7.
This clearly demonstrated that at lower heating rates ( 30°C/min)
two specific exotherms occurred associated with the individual decom-
position of OBSH and AC1. However, at all heating rates above
30 ° C/min the exotherms merged (as in Figure 6) which resulted in
rapid gas evolution over a narrow temperature range. This defeated the
object of this exercise, since the heating rate use in foaming studies was
in excess of 30 ° C/min. Tb obtain a CBAS of more universal applicabil-
ity, it was necessary to engineer the system with a view to maintaining
gradual gas release over a wider range of heating rates. In order to
achieve this, the OBSH concentration was reduced further (to reduce
the associated exotherm) and ZnO levels were increased to bring the
AC 1 decomposition temperature into the typical foam processing
range. Figure 8 shows the effect of two further formulations which took
the above into account.
Gradual gas release was more pronounced in the AC1 100/OBSH

Figure 6. DSC curve for mixture AC1 100/OBSH 50/ZnO 2 at 30°C/min.


A CBAS Based on Azodicarbonamide 183

Figure 7. The effect of heating rate on decomposition temperatures of the individual


blowing agent components in the CBAS: AC1 100/OBSH 50/ZnO 2 (parts by weight,
sample size = 0.010 g).

Figure 8. Decomposition behaviour of alternative formulations (parts by weight, sam-

ple size = 0.010 g).


184 G. L. A. SIMS AND H. A. S. JAAFAR

25/ZnO 5 system. Conversely, the formulation AC1 100/OBSH 30/ZnO


5 showed that the OBSH exotherm and ZnO activation influence was
still too great resulting in premature ADC decomposition which
adversely affected the rate of gas release, i. e., too rapid. Consequently,
the preferred formulation, AC1 100/OBSH 25/ZnO 5, was investigated
further as a function of heating rate by DSC. Figure 9 showed that the
decomposition temperatures remained separated up to a critical heat-
ing rate of 100 ° C/min. As the compression moulding technique used for
comparative foaming investigations did not exceed this critical rate,
the preferred AC1 100/OBSH 25/ZnO 5 system was used and compared
to AC1 100/ZnO 5 (activated AC 1 ). Micrographs of these foams are
shown in Figures 10 and 11.
Foaming with activated AC1 showed an irregular cell structure
(Figure 10) with a wide cell size distribution and evidence of small par-
ticulate decomposition residues seen within some of the cells.
Conversely, expansion using AC1 100/OBSH 25/ZnO 2 (which had
been shown to produce more gradual gas evolution) produced a sig-
nificantly improved regularity of cell structure (Figure 11) and no ap-
parent particulate residues.
Another interesting point is the consideration of the gas evolution at
6 ° C/min of the formulations AC 1 100/OBSH 50/ZnO 2 (Figure 4) and
the alternative preferred formulation AC1 100/OBSH 25/ZnO (Figure
8). Both curves bear some similarity (but it has been shown that only

Figure 9. The influence of heating rate on decomposition temperatures of individual


blowing agent components of the preferred formulation: AC 1 100/OBSH 25/ZnO 5 (parts
by weight, sample size== 0.010 g).
A CBAS Based on Azodicarbonamide 185

Figure 10. Micrograph of EVA foam specimen expanded using activated AC1.

Figure 11. Micrograph of EVA foam specimen expanded using AC1 100/OBSH
25/ZnO 2.
186 G. L. A. SIMS AND H. A. S. JAAFAR

the latter system can be used over a wide range of heating rates). This
is further substantiated by corresponding DSC measurements (Figures
5 and 12). Although the temperatures of the peak exotherms are identi-
cal within the limits of experimental error, it is noticeable that the
OBSH decomposition exotherm in the preferred formulation (AC1
100/OBSH 25/ZnO 5) is significantly reduced. On the basis of this work,
it appears that the CBAS requires a proportion of auxiliary BA high
enough to produce a limited initial exotherm to kick the ADC.
The magnitude of the auxiliary BA exotherm appears critical if it is
too high. Activation of the main blowing agent can occur too swiftly
resulting in decomposition of both blowing agents simultaneously (ex-
emplified by AC1 100/OBSH 100 in Figure 3). The findings of this paper
therefore point to the need for a delicate balance of the overall system.
It has also been shown that as heating rate affects the decomposition of
each of the constituent BAs, the CBAS must also be balanced to the pro-
cess heating rate. The preferred formulation in this work addresses the
above arguments but has not been optimized. It is stressed that all gas
evolution measurements were carried out in the absence of polymer
matrix, and all comparisons were based on the same sample weights.

Figure 12. DSC curve for mixture AC1 100/OBSH 25/ZnO 5 at 6°C/min.
A CBAS Based on Azodicarbonamide 187

The decomposition behaviour may be affected once dispersed in a


polymer. However, foaming studies indicated the general validity of
assumptions made.
CONCLUSIONS

It is possible to modify decomposition behaviour of ADC by using an


auxiliary BA. On the basis of this limited work, it appears that the
more gradual the gas release the more uniform the cell structure.
Selection of an appropriate auxiliary BA with a decomposition exo-
therm at a temperature sufficiently close to the decomposition temper-
ature of the main BA can activate the decomposition reaction of the
main BA. To obtain gradual release it is necessary to tailor the proper-
ties of the auxiliary and main blowing agents and main BA activator to
retain major decomposition performance at separate temperatures. It
appears that formulations must be balanced in terms of the magnitude
of the auxiliary BA exotherm. Process heating rate must also be taken
into account.

ACKNOWLEDGEMENT

The authors wish to express their appreciation to Mr. Raman Puri for
his helpful and supportive discussions and Schering Polymer Additives
for the supply of raw materials and use of certain instruments. One of
the authors (GLAS) also is grateful for support from Courtaulds plc. in
the form of the Courtaulds Lectureship in Polymer Technology.

REFERENCES

1. Nass, L. I. 1963. Modern Plastics, 40:151-237.


2. Puri, R. R. and K. T. Collington. 1988. Cellular Polymers, 7:56-84.
3. Jaafar, H. A. S. and G. L. A. Sims. 1993. Cellular Polymers
, 12:303-316.
4. Shutov, F. A. 1991. Foamed Polymers Processing and Production Technology
,
Basel, Switzerland, Technomic Publishing AG.
5. Klempner, D. and K. C. Frisch, eds., 1991. Handbook of Polymeric Foams
and Foam Technology, New York: Hanser Publishers.
6. Puri, R. R. and K. T. Collington. 1988. Cellular Polymers
, 7:219-231.

BIOGRAPHIES
H. Jaafar

Haydar Jaafar obtained a B.Sc. degree in Physics from Baghdad


188 G. L. A. SIMS AND H. A. S. JAAFAR

University, Baghdad 1987, and a M.Sc. degree in Polymer Science and


Technology from UMIST, Manchester 1993. He has worked on the be-
haviour of azodicarbonamide chemical blowing agent. He has trained
and worked on blowing agent systems performance in polyolefin and
PVC foams in conjunction with Schering Polymer Additives, Widnes.
He is now working as a visitor at UMIST Materials Science Centre
with Dr. G. L. A. Sims.
G. L. A. Sims

G. L. A. Sims graduated with a lst Class (Honours) B.Sc. degree in


Polymer Technology in 1969 and a Ph.D. in 1972 from UMIST. During
his early industrial career, he worked on the development and
manufacture of cross-linked PE foams for the German company Carl
Freudenberg, Weinheim. In 1980 he joined Jiffy Packaging Co. Ltd. and
became Technical and Works Director with particular responsibilities
for the development of noncross-linked PE foam.
In 1988 the company was acquired by the Dutch multi-national
Buhrmann-Tetterode and he was made Director of R&D of their Flexi-
ble and Protective Packaging Division. His prime task was to coor-
dinate and control development efforts in Europe and USA to ensure
implementation of CFC free foam production by 1990.
In late 1990 he returned to UMIST Materials Science Centre as
Courtaulds Lecturer in Polymer Technology. Major research interests
include blowing agent systems and impact behaviour of cellular
polymers.

Anda mungkin juga menyukai