Вы находитесь на странице: 1из 7

J Polym Res (2012) 19:9869

DOI 10.1007/s10965-012-9869-6

ORIGINAL PAPER

Accelerated thermal ageing studies of polydimethylsiloxane


(PDMS) rubber
Kewei Xiang & Guangsu Huang & Jing Zheng &
Xiaoan Wang & Guang xian Li & Jingyun Huang

Received: 10 January 2012 / Accepted: 3 April 2012 / Published online: 25 April 2012
# Springer Science+Business Media B.V. 2012

Abstract Thermal ageing studies over relative high temperature range (from 250 C to 85 C) have been conducted on
polydimethylsiloxane (PDMS) rubber with the aim of investigating ageing behavior. Tensile elongation, compression set
and creep measurements are introduced to monitor the ageing
process. Time temperature superposition is performed and the
linear trends are found in the Arrhenius plots for all the
approaches, revealing an identical process dominates the degradation process. Activation energies are found to be
depended on the methods used, which are 88.5 kJ/mol for
tensile tests, 67.7 kJ/mol for compression set measurements
and 75.3 kJ/mol for creep measurements. The phenomenon
that activation energy obtained from tensile elongation is
always higher than other approaches is rationalized by the
sensitive modification of thermal degradation on the rubbery
network. Moreover, correlations between the results from
different approaches have been examined. The linear relationships have been found between tensile elongation and creep or
compression set measurements for the three methods are all
influenced by the network structure.
Keywords Polydimethylsiloxane . Thermal ageing .
Lifetime estimation . Time temperature superposition

Introduction
Thermal ageing mechanisms for polydimethylsiloxane
Thermal degradation of polymeric materials have long been of
great interest both in academic and industrial perspective [13].
K. Xiang : G. Huang (*) : J. Zheng : X. Wang : G. xian Li :
J. Huang
College of Polymer Science and Engineering, State Key
Laboratory of Polymer Materials Engineering, Sichuan University,
Chengdu 610065, China
e-mail: guangsu-huang@hotmail.com

As a non-carbon backboned polymer, silicone rubber has been


widely used as damping material and sealing o-rings due to the
excellent performance on thermal resistance,non-stick, non
adhesive properties and low toxicity. Among the above advantages, the key feature of this material is the excellent thermal
resistance compared to traditional carbon backboned polymers,
which may also in principle result in a different degradation
mechanism. As the first member of this series of polymers,
polydimethylsiloxane (PDMS) owns the simplest chemical
structure and has been well studied in intensive literatures.
Similar to other polymers, PDMS tends to degrade when
exposed to ageing conditions like thermal, humidity and
radiation environments. A number of researches have been
accumulated in studying the ageing mechanism. The most
comprehensive work carried out by Grassie and co-workers
[4] illustrated that PDMS tends to degrade in a stepwise
fashion at elevated temperature. Degradation products contain volatile cyclic oligomers which mainly involve trimer
siloxane and little amount of other components from D3 to
D7 [5], residue contributions are chemically identical to
original polymer. The overall degradation pathways are
suggested to be backbiting [4, 5] for hydroxyl terminated
PDMS and intramolecular depolymerization for endblocking polymers [6] shown in Scheme. 1, which is
evidenced by the experimental findings that the decrease
in molecular weight is in proportional to the amount of
volatile materials, otherwise the molecular weight would
decrease sharply for random scission reactions [4, 7]. In
addition, the degradation process has been reported to be
easily catalyzed by both Lewis acid and hydroxyl ions
[810]. The former mainly involves -ion oxide which is
recognized to be able to scavenge free radicals and is widely
used as antioxidant in rubber industry. The later includes
some inorganic alkali such as KOH. The hydroxyl ions
would consequently result in siloxy ions, exerting major
influence on the oligomer forming reaction and accelerating
degradation process.

Page 2 of 7

J Polym Res (2012) 19:9869

Scheme 1 Thermal
degradation pathways for
PDMS. Key: a degradation
process for hydroxyl terminated
PDMS; b degradation process
for end blocking PDMS

Lifetime prediction for thermal ageing process

Experimental

Concerning thermal ageing, the most widely used method to


predict the lifetime of polymeric systems is the extrapolation
of failure time under accelerated conditions to room temperature, known as Dakins method [1113]:
 
ln "
E
ln tf ln

1
A
RT

Materials

Where is the normalized properties of interest such as


tensile elongation, compression set, induction time and so
forth, tf is the failure time, A is the preexponential factor, E
is the apparent activation energy, T is the absolute temperature
and R is the gas constant. This equation is suitable when there
is only one reaction cause the degradation, otherwise requiring
modified type [1418]. However, we essentially make use of
shift factor (T) to account for the temperature dependence of
a degradation process [1931]. This is because the shift factor
can be easily normalized to any temperature desired (i.e. T 01
at the lowest temperature or at the room temperature) and this
method could make use of data more efficiently. The relationship between T, tf and the rate constant (k) of a degradation
process are listed as follows [14]:
k

1
/ aT
tf

As a kind of commonly used polymer, although many


efforts have been driven on the investigation of the ageing
mechanism of PDMS rubber, less works have been devoted to
the activation energy at relative high temperatures. Moreover,
traditional mechanical tests have always been used separately,
the inherent correlations with other approaches need to be
further examined. In this work, some methods are introduced
to make the estimation through time-temperature superposition and the results were found to be verified by each other.

PDMS gum studied in this work was purchased from Chenguang Chemical Corporation, Si Chuan China, with a weightaverage molecular weight Mw of 580 kg/mol (Mw/Mn 02.3, as
determined by means of gel permeation chromatography using polystyrene as standard). For the poor mechanical performance of silicone rubber in its native state, fumed silica
(Chen-guang Chemical Corporation, BET surface area is
160 m2/g) is added into the bulk with the aim of improving
mechanical properties. Rubber blends were prepared in an
open two-roll mixer, and were homogenized in the following
recipe: 100phr PDMS, 2.5phr 2, 5-Dimethyl-2, 5-Di (tertbutyl peroxide) hexane (DPBMH), 40phr ultra fine silica. This
blend was press cured under 10 MPa at 170 C for 10 min and
was post-cured at 180 C for 4 h.
Oven aging
Accelerated aging experiments were carried out using an
air-circulating oven (accurate to 1 C) with stainless grids
located in it to support samples. Temperature ranges chose
for accelerated aging were 250 C to 90 C for tensile tests,
210 C to 85 C for compression set measurements and
250 C to 90 C for creep measurements. The compression
set values decreased so fast at 250 C within a few hours
that we limited our ageing temperatures for compression set
measurements to below 250 C.
Tensile tests
Tensile tests were performed on an Instron 5567 machine
(Instron, America) equipped with pneumatic grips and

J Polym Res (2012) 19:9869

extensometers clamped on the sample. Strain rate of this test


was 500 mm/min, five repeat samples under the same ageing condition (temperature and time) were typically measured and averaged to reduce errors.

Page 3 of 7
Table 1 Relationship
between the attached
weights and shore A
hardness of samples

Shore A hardness

35~44

4.3

45~54

5.4

55~64

Compression set measurements


Samples used for compressions set measurements were cylindrical (290.5 mm (diameter)*12.50.5 mm (height)).
This measurement was performed by first measuring the
height of unaged samples as h0. Then the strain height of
samples in the compression set jig was measured as hs
(9.34 mm). Samples were thermally aged under compression state in the jig placed in the oven. After removing
samples from the jig and the oven, cooling down at room
temperature for 30 min, the recovery height was measured
immediately as hr. Finally the compression set was calculated as h0  hr =h0  hs .
Creep measurements
Creep measurements were carried out using an air circulated
oven equipped with lever pressure devices (schematic diagram was shown in Fig. 1). A sample with the dimension
similar to the compression set measurements was pressured
by a weight hanging to the end of the lever to produce the
initial deformation of 20 2 %. And the choice of this
pressure was dependent on the Shore A hardness of samples. A correlation between the pressure and the hardness
was given in Table 1. For the hardness of the investigated
sample was 56, the attached weight was chosen as 7.5 kg.
From the lever device, samples underwent compression
force tenfold larger than the essential weight of the attached
briquette. The creep values during thermal aging were measured using a thickness gauge with the accuracy of
0.01 mm.

Fig. 1 Schematic representation of the creep instrument

Attached
weight (kg)

7.5

65~75
76~85

12
33

86~95

73

Results and discussion


Time-temperature superposition and the Arrhenius analysis
Samples used for tensile tests were aged at nine temperatures and the results were displayed in Fig. 2. At each
temperature, about ten data points were recorded to draw
the overall aging profiles. It could be seen that the tensile
elongation decreased gradually with ageing time. Unlike
many other polymers such as EPR rubber [31], the induction behavior (properties changed slowly until just before
failure) did not appear. It is well known that the most
important principle for time-temperature superposition is
based on the stress-independent damage model [32],
which means that the shapes of the curves in the whole
temperature range do not change and are independent of
the degradation stress level, otherwise they can not be
superposed well by simply shift curves horizontally at elevated temperature to the reference one. From Fig. 2, tensile
elongation results under different temperatures had the similar shapes and thus showed the feasibility of superposition.
In this work, we chose the lowest temperature (90 C) as the
reference one.

Fig. 2 Normalized tensile elongation results versus aging time at nine


temperatures, the error bars represent the standard deviation

Page 4 of 7

As mentioned above, theories of time-temperature superposition essentially base on the thought that thermal degradation rate constants k increases as the raising temperature.
Quantitatively, a plot of ln(k) versus inverse absolute temperature (1/T) would yield a straight line and the activation
energy could be calculated from the slope. Figure 3 showed
the extrapolation profiles of the temperature range investigated; the shift factors were shown in the internal figure.
According to the superposition, the master curve was
obtained by horizontally moving higher temperature curves
to the reference one using a constant scaling factor (aT) to
give the best superposition. And the relationship between
the derived lnaT versus reciprocal absolute temperature was
plotted in Fig. 4.
The data sets in Fig. 4 appeared to show the Arrhenius
behavior, with no curvature points along the straight line.
From the slope (10.645) of the line, activation energy could
be calculated as E010.6458.314088.5 kJ/mol, which was
consistent with the results from M.Patel [10] on the investigation of room temperature vulcanized silicone rubber (77
45 kJ/mol for temperature above 150 C). In that work, the
extrapolation curve turned to lower activation energy (22
7 kJ/mol) below 130 C, indicating that 130 C was a
temperature where the underlying degradation process
changed. The curvature was rationalized by the acid accelerated hydrolysis due to the similar activation energy [33].
However, from Fig. 4 it could be clearly seen that no
curvature emerged around 130 C. To confirm the results
of this extrapolation, other approaches were introduced.
Figure 5 displayed the results of compression set measurements with the temperature ranging from 210 C to 85 C.
The compression set values display fairly regular changes
compared to tensile elongation. Also, time-temperature superposition was performed. The superposition profile was
given in Fig. 6 along with the Arrhenius plot in Fig. 7. By

Fig. 3 The master curve of tensile elongation tests with 90 C as the


reference temperature, the shift factors are as indicated

J Polym Res (2012) 19:9869

Fig. 4 Arrhenius plot of shift factors and the extrapolation results at


room temperature

analogous with the tensile results, logarithm form of the


master curve was chose for clarity. It could be clearly
observed that there was also no bending in the Arrhenius
plot and a linear type was evidenced, with the activation
energy calculated as E08.1398.314067.7 kJ/mol. This
value presents a decrement than that from tensile results,
suggesting a weaker temperature dependence of compression set measurements.
As another condition monitoring method, creep measurements were also performed in the temperature range of 250 C
to 90 C. Figure 8 shows the isotherms under different temperatures. Similar superposition procedure was carried out with
the master curve displayed in Fig. 9. Excellent overlap was
obtained and the shift factors were listed with corresponding
temperature in the internal figure. The Arrhenius plot in
Fig. 10 also verified no slope change in the investigated
temperature range, which further confirmed the linear

Fig. 5 Compression set results versus ageing time with temperature


ranging from 210 C to 85 C

J Polym Res (2012) 19:9869

Fig. 6 The master curve of compression set measurements with 85 C


as the reference temperature, the shift factors are as indicated

Page 5 of 7

Fig. 8 Isotherms of creep measurements

extrapolation. From the slope of the straight line, activation


energy of creep measurement was calculated as E09.06
8.314075.3 kJ/mol, this value is intermediate between the
results from compression set and tensile elongation tests.
From the above results, it could be seen that activation
energies are dependent on the methods used. Many researches
[22, 2527] in the study of thermal ageing of various polymeric
materials illustrated that activation energy deduced from tensile
elongation tests is larger than those from other approaches. In
other words, tensile elongation test is the most sensitive to
thermal ageing and has the strongest temperature dependence.
Generally speaking, mechanical properties of polymers will be
influenced by the following factors: polydispersity, average
molecular weight between two crosslinking points, the concentration of elastically active chains and the intermolecular interactions such as hydrogen bonding [34, 35]. For the
intermolecular interactions of a specific polymeric system are

identical, thus low polydispersity, long molecular chains between crosslinking points and a perfect network with less
dangling ends will result in better mechanical performance.
When samples undergo thermal ageing, the crosslinking network and chemical construction of backbones will be modified
greatly. For silicone rubber, the degradation process is dominated by the depolymerization and chain scission reaction,
leading to the cleavage of mainchain and producing amounts
of dangling ends which may be the weak points on applying
strain to samples. Moreover, it must be noted that the correlation between the tensile elongation and the modification on the
molecular level is very subtle despite no quantitative model was
found in the literature. A little change in the chemical structure
and the crosslinking network will exert a great influence on the
mechanical properties, making tensile elongation much more
sensitive to the effect produced by thermal ageing, which is the
reason why this approach has the strongest temperature dependence compared to other methods.

Fig. 7 Arrhenius plot and the extrapolation line of compression set


measurements

Fig. 9 The master curve of creep measurements with 90 C as the


reference temperature, the shift factors are as indicated

Page 6 of 7

J Polym Res (2012) 19:9869

Fig. 10 Arrhenius plot and the extrapolation profiles of creep


measurements

Fig. 12 Correlation between the results from tensile elongation and


creep measurements

Extrapolation profiles and the correlations between three


approaches
According to the Arrhenius theory, activation energy of the
whole aging process and curves of mechanical properties
versus aging time at any desired temperature could be derived
from extrapolation. With respect to tensile tests, shift factor at
room temperature could be calculated as aT exp6:64
1=765 as shown in Fig. 4, then the logarithm plot of
normalized tensile elongation at room temperature was
obtained by multiply the x axis of curve at 90 C by 765.
The corresponding plot was shown in Fig. 11. Lifetime could
be estimated if the failure criterion is ascertained. However,
we need to verify the feasibility of linear extrapolation on the
Arrhenius plot over wide temperature range. Expanding the
ageing experiments to lower temperature (such as 60 C) to
compare the prediction values and the thermal ageing results

would require lengthy ageing time and become out of traditional time scale (which may up to dozens of years). This work
has still been carried out in our laboratory and on the other
hand, more sensitive analytical methods need to be developed.
However, it is worth noting that lifetime of silicone rubber
under field ageing can not reach the expected lifetime from the
Arrhenius extrapolation. This is because silicone rubber is
driven to degrade via many other combination effects such
as irradiation or humidity, not only thermal oxidation. The
humidity may be the dominating factor causing degradation at
lower temperature. Consequently lifetime expectancy based
on time-temperature superposition should be carefully applied
to accelerate thermal ageing under complex degradation
environment.
The inherent relationships between results from creep
and compression measurements with elongation need to be
established. From the correlation plot shown in Figs. 12 and

Fig. 11 Tensile elongation versus aging time at 25 C obtained by


extrapolation of master curve

Fig. 13 Correlation between the results from tensile elongation and


compression set measurements

J Polym Res (2012) 19:9869

13, it could be seen that when the normalized tensile elongation reached the arbitrary selected failure point (0.5), the
creep value would approach 0.48 mm, and the compression
set value came close to 0.62 despite a little deviation of data
points at 130 C. In addition, the linear relationships are
identified in both of the two figures, this is because all the
methodologies used in this work were strongly influenced
by chemical crosslinking networks. The linear relationships
between tensile elongation and other approaches have been
observed by some other authors such as elongation versus
NMR transverse relaxation time (T2), solvent uptake factor,
gel percentage and etc. [36, 37]. As stated above, this is due
to the changes in rubbery network lead to the modification
both on macroscopic and microscopic properties which are
detected by different analytical methods.

Conclusion
Thermal ageing process of PDMS rubber was studied using
tensile elongation, creep and compression set measurements.
The activation energies derived from time-temperature superposition were found to be dependent on the methodologies
used, which were 88.5 kJ/mol for elongation, 75.3 kJ/mol for
creep measurements and 67.7 kJ/mol for compression set
measurements. The strongest temperature dependence of tensile elongation among other approaches was illustrated by the
easily modification of thermal ageing on the rubbery crosslinking network. Through correlating the results from different
approaches, linear relationships between tensile elongation
and creep or compression set measurements were demonstrated. This was rationalized by the modification on the network
structure induced by thermal oxidation.
Acknowledgments The authors appreciate the financially support
from the National Natural Science Foundation of China (Grant No.
51133005).

References
1. Kaya , Avc A (2012) J Polym Res 19:97809794
2. Ramazani S, Morshed M, Ghane M (2011) J Polym Res 18:781793
3. Jin J, Chen S, Zhang J (2010) J Polym Res 17:827836

Page 7 of 7
4. Grassie N, MacFarlane IG (1978) Eur Polym J 14:875884
5. Lewicki JP, Liggat JJ, Patel M (2009) Polym Degrad Stab
94:15481557
6. Thomas TH, Kendrick TC (1969) J Polym Sci A Polym Chem
27:537549
7. Grassie N (1966) Encyclopedia of polymer science and technology, vol 4. Wiley, New York, p 647
8. Dean PR, Kuczkowski JA (1985) Rubber Chem Technol 59:842
9. Yang ACM (1994) Polymer 35:32063211
10. Patel M, Skinner AR (2001) Polym Degrad Stab 73:399402
11. Dakin TW (1960) Electra-Technol 66:123
12. Dakin TW (1984) Trans AIEE 67:113
13. IEC Standard 216 (The corresponding British Standards are BS56911995): Guide for the Determination of Thermal Endurance
Properties of Electrical Insulating Materials. Bureau Central de la
CEI, Geneva, Switzerland
14. Celina M, Gillen KT, Assink RA (2005) Polym Degrad Stab
90:395404
15. Brown RP (2001) Practical guide to the assessment of the useful
life of rubbers. Rapra Technology Limited, Shawbury
16. Budrugeac P, Segal E (1994) Polym Degrad Stab 46:203210
17. Budrugeac P (1995) Polym Degrad Stab 50:241246
18. Gugumus F (1999) Polym Degrad Stab 63:4152
19. Morrell PR, Patel M, Skinner AR (2003) Polym Test 22:651656
20. Gillen KT, Bernstein R, Clough RL, Celina M (2006) Polym
Degrad Stab 91:21462156
21. Bernstein R, Gillen KT (2009) Polym Degrad Stab 94:2107
2113
22. Wise J, Gillen KT, Clough RL (1995) Polym Degrad Stab 49:403
418
23. Celina M, Wise J, Ottesen DK, Gillen KT, Clough RL (2000)
Polym Degrad Stab 68:171184
24. Celina M, Minier L, Assink RA (2002) Thermochimica acta
384:343349
25. Gillen KT, Bernstein R, Derzon DK (2005) Polym Degrad Stab
87:5767
26. Dong WF, Gijsman P (2010) Polym Degrad Stab 95:10541062
27. Gillen KT, Bernstein R, Celina M (2005) Polym Degrad Stab
87:335346
28. Bernstein R, Gillen KT (2010) Polym Degrad Stab 95:14711479
29. Denardin ELG, Janissek PR, Samios D (2003) Thermochimica
acta 395:159167
30. Gillen KT, Celina M, Bernstein R (2003) Polym Degrad Stab
82:2535
31. Gillen KT, Celina M, Bernstein R, Shedd M (2006) Polym Degrad
Stab 91:31973207
32. Hwang W, Han KS (1986) J Compos Mater 20:125153
33. Stein J, Prutzman L (1988) J Appl Polym Sci 36:511521
34. Colin X, Audouin L, Verdu J (2007) Polym Degrad Stab 92:906914
35. Llorente MA, Mark JE (1980) Macromolecules 13:681685
36. Gillen KT, Assink RA, Bernstein R (2004) Polym Degrad Stab
84:419431
37. Gillen KT, Assink RA, Bernstein R, Celina M (2006) Polym
Degrad Stab 91:12731288

Вам также может понравиться