You are on page 1of 8

Polymer Degradation and Stability 41 (1993) 109-116

Effect of compatibiliser, curing sequence and ageing on the thermal stability of silicone rubber, E P D M rubber and their blends
S. Kole, T. K. Chaki, Anil K. Bhowmick & D. K. Tripathy*
Rubber Technology Centre, Indian Institute of Technology, Kharagpur 721 302, India
(Received 17 July 1992; accepted 6 August 1992)

The thermal stability of the vulcanisates of silicone rubber, EPDM, their blends and aged specimens of these systems has been studied by nonisothermal thermogravimetry. The activation energy for the degradation process suggests that a different route is followed for the degradation of EPDM after ageing. The effect of a silane-grafted EPR and also the effect of a programmed two-stage curing process on thermal stability has also been studied. Both are found to contribute moderately to the thermal stability of the blends.

INTRODUCTION
The stability of rubber vulcanisates depends largely upon the environment, because degradation is primarily dependent upon thermal or thermal oxidative degradation, or both. Thermal stability becomes an important factor when the polymer is highly resistant to oxidation, or when it is heated in an inert atmosphere, or when the rubber is in the form of a thick section, so that oxidation becomes diffusion-controlled. The structure of the main chain, the energy of the main chain bonds, the nature of the crosslinks and the presence of any extra-network material, all affect the thermal stability of rubber vulcanisates. Thermogravimetric analysis, together with other supporting experiments like efffluent-gas analysis, IR and X-ray diffraction can all help in understanding the degradation mechanism and must assist any effort to enhance the thermal stability of a polymeric material.~ 3 The thermal stability of polysiloxane elastomers is ascribed to the high strength (110 kcal) of its main chain bonds (Si--O). 4 The thermal
* To whom correspondence should be addressed.

Polymer Degradation and Stability 0141-3910/93/$06.00 1993 Elsevier Science Publishers Ltd.
109

stability of E P D M is attributed to its saturated main chain structure. 5 As far as high-temperature stability is concerned, silicone rubber is superior to EPDM, 6 but the insulation characteristics of E P D M seem to be better. 6'7 Thus, a suitable blend of silicone-EPDM can give a product with generally superior properties to either polymer. However, silicones are known to be incompatible, in general, with other polymers. 8 This explains why simple blending and subsequent vulcanisation do not yield a material with improved properties. The present work aims at improving the heat-resistance properties of silicone-EPDM blends, either by the use of a compatibiliser or a programmed sequence of mixing and vulcanisation. The effect of ageing on the thermal stability of the blends has also been studied. Sometimes, specifically when a fire breaks out, polymeric materials undergo degradation in a restricted supply of oxygen or in a closed chamber, and the entire process can be visualised as an initial thermal oxidation followed by subsequent thermal degradation. Such a degradation process is gradually attracting more attention because it generates mt;ch smoke and toxic gases which endanger lives. The controlled laboratory simulation of such a degradation sequence can be represented best by initial ageing of the specimens in a restricted supply of

110

S. Kole e t al.

oxygen, followed by thermal degradation in an inert atmosphere.

EXPERIMENTAL Materials
Silicone rubber. Silastic 1625 (Dow Corning), type VMQ (i.e., vinyl-methyl-based silicone), specific gravity 1.25 (25C), and E P D M rubber. Keltan 520 [ethylene content 55 mole %, diene content 4.5 mole % (DCPD), density 0.86 g/cm 3] were both supplied by Fort Gloster Industries Ltd., India. Compatibiliser. Silane-grafted E P R was developed in our laboratory following the procedure of Sen et al.9 Peroxide. Dicumyl peroxide was procured from Hercules Incorporated, USA.

was blended with the desired amount of E P D M in the Brabender plasticorder at the same temperature and speed for 5 min as before. In the single-stage curative mixing process as generally followed, the curative was mixed in a cold two-roll mill. In a few cases, part of the curative was mixed in the Brabender at 120C and at a rotor speed of 100rpm for various periods of time to ensure a certain degree of cure, and the balance at room temperature in the roll mill. Curing was carried out in a hydraulic press at a pressure of 5 MPa at 170C for 10 min. Post curing was done at 150C for 2 h in a temperature-controlled air-circulated ageing oven (model No. FC 712, Blue M. Electric Co., Blue Island, Illinois, USA).

Agemg
Ageing was carried out in a Test Tube Ageing Tester (Seisaku-SHO, Toyoseiki, Japan) at 175C for 72 h.

Mixing and vulcanization


Table 1 shows the composition of the mixes used in the present investigation. Blending of silicone rubber with E P D M was carried out in a Brabender plasticorder (PLE 330) at 120C for 5 min at a rotor speed of 100 rpm. Compatibiliser was mixed, whenever applicable, with silicone rubber in the plasticorder under identical conditions, and the masterbatch thus obtained

Thermogravimetric analysis
Thermogravimetric analysis was done in a 951 Thermogravimetric Analyser fitted to a 9000 model Thermal Analyser from DuPont. Thermal degradation of aged and unaged vulcanisates was studied non-isothermally from ambient temperature to 800C at a programmed rate of 20C/min

Table 1. Compositions of the mixes


Mix Silicone (parts) EPDM (parts) Silanegrafted EPR (parts) -----8 10 15 -DCP (parts)

A B C D E F G H I J K

100 75 50 25 -50 5O 50 50 50 50

-25 50 75 100 50 50 50 50 50 50

1.5 1.5 1.5 1.5 1.5 1.5 1.5 1.5 0-5+1

(1)
0.25 + 1.25

(3)
0.13 + 1-37

(3)
Samples I, J and K were cured in a two-stage process as described in the experimental section. The second n u m b e r under D C P indicates the a m o u n t added in the second stage of mixing. The figures in parenthesis indicate the curative mixing time (in mins) in the first stage.

Effect of compatibiliser, curing sequence and ageing

111

in a nitrogen atmosphere. Thermogravimetric and differential thermogravimetric curves were obtained from the plotter.

position step only. The order of the decomposition process for silicone rubber, EPDM and for the first decomposition step of the blends was found to be approximately unity.

Analysis of thermograms
Order and activation energy for the decomposition reaction were determined using the standard kinetic equation dw
- =

RESULTS AND DISCUSSION Degradation of the base elastomers


Figures 1 and 2 show the thermal degradation behaviour of the vulcanisates of silicone rubber (mix A), EPDM rubber (mix E) and a representative blend (mix C, 50/50 blend of silicone/EPDM). From these data, it is clear that both the pure vulcanisates undergo a one-step decomposition process. Further, silicone rubber decomposes over a wide temperature range and EPDM decomposes over a narrow temperature range. The thermal stability of silicone rubber is found to be superior to that of EPDM rubber as reflected by its significantly higher values of IPDT and Tmax (Table 2). However, the kinetic parameters apparently contradict this. The activation energy (Table 2) for decomposition of silicone rubber (25 kcal/mole) is lower than that of EPDM (78 kcal/mole) and the pre-exponential factor is very much lower, 105 compared with 1022 , which indicates relatively few degradation sites for silicone rubber compared to EPDM.

Ae-e/grw .

dt

where w denotes the weight fraction of the active material remaining, and A, E, T, n and t have their usual meanings. 1 The equation was solved graphically for E and n using the thermogravimetric (TG) and the differential thermogravimetric (DTG) curves as described by Freeman and Carroll. '1 The pre-exponential factor, A, was determined from the original equation using the known values of E and n, and the range of values obtained over the entire thermogram is reported. Integral procedural decomposition temperatures (IPDT) were determined from the TG curves in the usual way. Tmax,the temperature of the maximum decomposition rate, was read from the DTG curve. The degradation parameters, E, n and A, for the blends were calculated for the first decom-

~oo ~ ~ ~ x~ ~ . ..... -- . . . . .

EPDM (mix E) 5 0 1 5 0 blend ( m i x C) Silicone (mix A )

80

~ 6O
o

r:
4O

l ! \\\.\.
\
~'N%, N
%\ \ \ \

\.~.~

20

200

400 T e m p e r a t u r e ( C )

600

800

Fig. 1. T h e r m o g r a v i m e t r i c curves of silicone, E P D M and 50/50 blend of silicone and E P D M .

112 60

S. Kole et al.

..... ..... 40
r-

EPDM (mix E) 5050 blend (mix C) Silicone (mix A)

i/i o 0A

20 i t if I

*6
rr

" \,l.f~ .--. \"

-20 0

2o0

4oo Temperature (*C)

60o

8uo

Fig. 2. Differential thermogravimetric curves of silicone, EPDM and a 50/50 blend of silicone and EPDM.

This peculiarity in behaviour of the silicone vulcanisate can be u n d e r s t o o d if we recall the degradation mechanism of silicone which consists of oligomer elimination from the chain end. Thus, the actual degradation route, which determines the n u m b e r of available sites that can participate in the degradation process, is also very important in dictating the thermal stability of a specimen.
Degradation of the blends

Blends degrade in two steps (Figs 1 and 2 and Table 2). The first Tmaxvalue (Tmax,1) in the D T G curve coincides with the Tmaxvalue of E P D M and

the second Tmax value (Tm~,2) gradually shifts to lower temperatures from the Tm~ value of silicone as the E P D M concentration is increased. A t the peak decomposition t e m p e r a t u r e of E P D M (488C), silicone decomposes at a rate of 2.65% per min c o m p a r e d with a value of 55% per min for E P D M . Thus in the absence of strong specific interactions between the components, which is true for an incompatible system like s i l i c o n e - E P D M , the peak decomposition rate and t e m p e r a t u r e for the first decomposition will be controlled by E P D M so long as its concentration does not fall drastically. The detailed morphological study of s i l i c o n e - E P D M blends clearly confirms the incompatible nature

Table 2. Degradation behaviour of unaged specimens of silicone, EPDM and their blends

Mix

Pre-exponential factor (1.9-2.9) x 105 (1-2-3-3) x 10'2 (1-1-3.6) x 1015 (2-3-5) 1015 (1.4-2.7) x 1022

Activation energy (kcai/mole) 25 46 55 55 78

IPDT (C) 547 + 1 505 + 1 490 -t- 2 485 + 1 485 + 1

Tm,x.ta (C) -488 + 2 488 + 2 488 + 2 488 + 2

T,~ax.2 a (C) 581 + 1 575 + 2 560 + 1 545 + 2 --

A B C D E

" Tm,x.~= T,,~ for EPDM or for 1st decomposition of the blends. Tin,x.2= Tmaxfor silicone or for 2nd decomposition of the blends.

Effect of compatibiliser, curing sequence and ageing


of the blend, and this is in line with the theoretical considerations of Sanchez on miscibility. 12'8 The second decomposition zone is possibly interfered with by the last trace of EPDM remaining and this explains the gradual shift of Tm~.2 towards lower temperatures with E P D M loading. Both the elastomers are cured with DCP, and curing proceeds through free radicals. Thus at the interface of the two polymers, some inter-crosslinking can occur, which may temporarily retain some of the E P D M fragments with the silicone even when the bulk of EPDM has volatilised. These inter-crosslinked fragments, at some elevated temperatures, will volatilise either by further fragmentation or by scission of the inter-crosslinks. This explanation seems to be very reasonable in view of the very low value of the peak decomposition rate of silicone and the correspondingly very high value of EPDM, so that retention of a little E P D M will be enough to modify the overall degradation rate in this zone. Values of IPDT, Tmax, and other kinetic parameters for silicone rubber, E P D M and their blends are shown in Table 2. Calculated values indicate the decomposition processes to be of first order. Values of E and A increase with EPDM concentration. Change in activation energy is simply a dilution effect and linear

113

change of activation energy with composition is also assumed by other workers. 13 IPDT is found to decrease with increase in E P D M content. Figure 3 shows the effect of ageing on thermal degradation behaviour. Results for aged specimens are shown in Table 3. The same trend is shown as in the case of unaged specimens. The main differences are reductions in activation energy, pre-exponential factor, IPDT and Tmax for E P D m and E P D M containing blends. The behaviour of silicone rubber does not alter with ageing. A lowering of the IPDT value indicates some loss in thermal stability on ageing. Reduction in activation energy implies that ageing offers an alternative degradation route. Reduction in pre-exponential factor suggests that new degradation sites are generated only at selected points on ageing. Thus it appears that ageing incorporates new potential degradation sites of lower activation energy and so leads to reduced thermal stability of E P D M and its blends.

Effect of compatibiliser
Figure 4 shows the effect of compatibiliser (mix F, 8 parts per hundred parts of rubber compatibiliser) on the thermal degradation

12(3 - ..... 10(3 EPDM ( u n a g e d m i x E ) EPDM ( a g e d mix E )

60

4O 80

~6c
40

2O _o

nO

20

-I I

. . . . . . . . . . . I

__.

200

400 T e m p e r a t u r e (C)

600

1-20 800

Fig. 3. TG and DTG curves of aged and unaged EPDM.

114

S. Kole et al.
Table 3. Degradation behaviour of the specimens of silicone, E P D M and their blends aged for 72 h at 175"C Mix Preexponential factor (1.2 x 4.2) x 106 (4.8 x 4.9) x 1 0 4 (1.8 x 2.3) x 106 (3.3-9) x 109 (0.9-3.4) X 1012 Activation energy (kcal/mole) 28 21 24 35 43 IPDT (C) 553 500 465 465 465 + + + + 1 1 2 2 2
T,....," 7",....~"

(C) -476 + 476 + 476 476

(C) 575 + 565 540 + 530 + -1 1 1 2

A B C D E

2 2 1 2

= See footnote to Table 2.

behaviour of a 50/50 silicon-EPDM blend. Results of degradation kinetics for specimens containing compatibiliser (mixes F, G and H) are shown in Table 4. Silane-grafted EPR acts as a physical compatibiliser that controls the domain sizes by reducing interracial energy. 12a4 As the interaction is present in the interface and the bulk of polymer remains uninfluenced, individual components are likely to follow their own degradation route. This explains why the compatibiliser does not improve the thermal stability significantly. In fact, at compatibiliser concentrations up to 10 phr, the improvement is only marginal and at 15 phr there is a decrease in this property. This reflects the dilution effect of EPR in the blend, whose thermal stability is identical to EPDM and inferior to silicone.
120

Apart from the interaction mechanism at the interface, the finer morphology generated by compatibiliser can lead to slower diffusion of reaction products by offering a more tortuous diffusion path owing to the physical barrier imposed by silicone to the diffusion of the volatiles. This can lead to a pseudo-stability of the blend.

Effect of restricting the domains by two-step curing


These specimens were prepared with a view to generating a co-continuous structure. During mixing, severe stratification of the phases and partial coalescence of the stratified structures can
30
..... ....... 50150 blend ( m i x C ) 50150 blend with 8 p hr compatibiliser ( m i x F) A t w o st age cur ed p r o d u c t (mix K)

13
100

20

80

~. 6C
c~

/,ii i!
lY

I
10

E
t~

~""
n~

2C

I 200

I 400 Temperature (*C)

I 600

I 8OO

-10

Fig. 4. T G and D T G curves o f a 50/50 blend, a specimen containing compatibiliser and a restricted-domain specimen.

Effect of compatibiliser, curing sequence and ageing


Table 4. Role of compatibiliser on thermal stability of silicone-EPDM blends
Mix Preexponential factor (1.1-3-6) x 10 ~5 (4-5-4) 1012 (5-9-4) x 109 (0.7-2-4) 10 L3 Activation energy (kcal/mole) 55 47 36 46 IPDT (C) 490 497 493 473 +2 :t: 2 5=2 3

115

Tm~x.,a
(C) 488 + 2 498 + 1 490 5:1 485 + 2

Tmax.2 a
(C) 560 568 560 538

C F G H

See footnote to Table 2.

Table 5. Effect of restricted-domain structures on thermal stability of silicone-EPDM blend


Mix PreActivation IPDT

T.~ax.,"

T.,ax.2a

exponential
factor C I J K (1.1-3-6) x 10 ~5 (2-1-3.9) x 1013 (2-3.6) x 10 t4 (3"7-8-4) x 10 ~

energy
(kcal/mole) 55 49 53 44

(C)
490 + 2 490 :t: 2 505:1:2 500-t-2

(C)
488 + 2 505 + 2 507 + 2 502+2

(C)
560 568 585 591

u See footnote to Table 2.

lead to a co-continuous structure, and it was thought that partial crosslinking of the phases during this stage might retain this morphology and also the continuous structure of the silicone phase would lead to improved heat ageing properties. ~2 Degradation behaviour and kinetics of restricted-domain specimens (mixes I, J and K) are shown in Fig. 4 and Table 5. The specimens do not show any significant change in activation energy. However, the pre-exponential factor decreases to some extent, whereas IPDT and Tmax,~show some improvement. This can be ascribed either to the difficulty of diffusion of the degradation products, or to some interfacial bonding between the two phases. During dynamic vulcanisation, a considerable amount of interface is generated and this ultimately enhances the possibility of inter-crosslinking. One surprising observation is that Tm~x,~for these specially prepared blends are higher than the T~,ax of silicone rubber, leading to the apparent conclusion that the process strengthens the silicone phase further.

(iii) Finer morphology generated by compatibiliser leads to higher thermal stability and this is attributed to two reasons, namely the difficulty of diffusion of the products and an inter-crosslinking reaction between the two elastomers. (iv) Restricted-domain specimens have higher thermal stability and the silicone phase seems to be more stabilised. (v) Partial oxidation can reduce the thermal stability drastically, since it creates points that are highly susceptible to degradation and provides an altogether different route for degradation.

REFERENCES
1. Stuckey, 2. 3. 4. 5. 6. 7. In Developments in Polymer eds. G. Scott. Applied Science Publishers, London, 1979. Grassie, N. (ed.), Developments in Polymer Degradation--1. Applied Science Publishers, London, UK, 1977. Kenyon, A. S., In Techniques and Methods of Polymer Evaluation, Vol. 1, eds P. E. Slade Jr & L. T. Jenkins. Marcel Dekker, New York, USA, 1966, p. 217. Yilgor, I. & McGrath, J. E. In Advances in Polymer Science-86, ed. H. Benoit, Springer Verlag, 1988, p. 6. Brydson, J. A., Rubber Chemistry. Applied Science Publishers, London, UK, 1978, p. 323. Technical Literature, JSR JENIX E. Japan Synthetic Rubber, Tokyo, Japan, 1990. Technical Literature, Royalene ~m EPDM for IM and RPO Applications. Uniroyal Chemical, Naugatuck, USA, 1989. J. E.,

Stabilisation--1,

CONCLUSIONS (i) Degradation of silicone, EPDM and their blends follow first-order kinetics. (ii) Degradation of the blends shows the effects of the degradation of the individual components.

116

S. Kole et al.
Interscience Publishers, New York, 1971, p. 7. 11. Freeman, E. S. & Carroll, B., J. Phys. Chem., 62 (1958) 394. 12. Kole, S., Bhattacharya, A. K. & Bhowmick, A. K., Piast. Rubb. Comp. Proc. Appln., 19 (1993) 117. 13. Feldman, D. & Rusu, M., Eur. Polym. J., 10 (1974) 41. 14. Kole, S., Bhattacharya, A. K., Tripathy, D. K. & Bhowmick, A. K., J. Appl. Polym. Sci. (in press).

8. Sanchez, I. C., In Polymer blends, Vol. 1, eds. D. R. Paul & N. Newmann. Academic Press, New York, 1978, Ch. 3. 9. Sen, A. K., Mukherjee, B., Bhattacharya, A. S., De, P. P. & Bhowmick, A. K., J. Appl. Polym. Sci., 44 (1992) 1153. 10. Reich, L. & Levi, D. W. In Encyclopedia of Polymer Science and Technology, Vol. 14, ed. N. M. Bikales.

You might also like