Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

International Journal of Adhesion and Adhesives 87 (2018) 64–72

Contents lists available at ScienceDirect

International Journal of Adhesion and Adhesives


journal homepage: www.elsevier.com/locate/ijadhadh

Thermomechanical behavior of aerospace-grade RTV (silicone adhesive)☆☆,☆ T


Todd M. Mower
M.I.T. Lincoln Laboratory, Lexington, MA 02420, USA

ARTICLE INFO ABSTRACT

Keywords: Measurements of mechanical properties of the commercial polysiloxane RTV 566 were conducted over a range of
Polyphenylmethylsiloxane temperatures, with particular focus on characterizing isothermal behavior at reduced temperatures to reveal
Low-temperature testing possible evidence of crystallization in this material. Conventional testing was performed using typical tem-
Thermal expansion perature sweeps to measure elastic flexural and tensile moduli, and thermal expansion coefficients with and
Elastic modulus
without volumetric constraint. The glass transition temperature was determined to be about −115 °C.
Shrinkage
Crystallization
Measurements of the tensile moduli of RTV 566 specimens were also conducted under isothermal conditions
at reduced temperatures. After exposure to temperatures below −75 °C for several hours, tensile moduli rose by
as much as a factor of 40. Volumetric shrinkage of 0.35% was demonstrated after 50 h under isothermal con-
dition at −80 °C. These measurements demonstrate previously unpublished manifestations of crystallization
occurring in this material, which is based upon a polydimethylsiloxane with phenyl substituents for some of the
methyl groups.

1. Introduction applications destined for deployment in space, chiefly because of its


low-outgassing behavior (meeting NASA requirements, as specified in
1.1. Motivation ASTM E595), low glass transition temperature (Tg), and enhanced
thermal conductivity. Despite widespread use of RTV 566 throughout
During environmental testing of a telescope lens assembly built at the aerospace and optical communities, very little data concerning its
MIT Lincoln Laboratory, the focus continued to shift after thermal thermomechanical behavior have been published in open literature.
equilibrium had been established at cold temperatures (−75 °C and Moreover, measurements of its mechanical properties have typically
−85 °C). This shift steadily developed increasing magnitude and sta- been conducted under ambient conditions or using linear ramps in
bilized after several days, and could be accounted for by neither ther- temperature. A primary goal of this study was to explore the tem-
moelastic mismatches nor viscoelastic behavior of the materials used in perature and strain conditions that might induce crystallization in RTV
the assembly. While considering other possible causes for the behavior 566.
of the lens assembly, crystallization of the silicone adhesive used to Compliant adhesives are very often used to bond brittle elements,
mount the lenses was suspected. Because very few studies of crystal- such as glass lenses, to stiff metal structures. Examples of the prominent
lization in silicone materials (typically polydimethylsiloxane, PDMS) use of RTV 566 within the aerospace community abound. In NASA's
have been published in the open literature, the investigation reported Geostationary Operational Environmental Satellite (GOES), RTV 566
here was undertaken. was used to bond the glass secondary mirror into a supporting structure
Room temperature vulcanizing (RTV) adhesives have been available made with Invar [2]. In the Geoscience Laser Altimeter System (built at
and in use for many decades. One very prominent RTV developed by NASA's Goddard Space Science Facility), RTV 566 was used to bond
General Electric over fifty years ago is based up PDMS, with phenyl optic mount shims in the Laser Select Mechanism. The Jet Propulsion
substituting for some of the methyl groups in the chemical configura- Laboratory has also used RTV 566 in critical assemblies built for NASA,
tion; this adhesive is marketed as RTV 566 and is classified as a poly- including the Mars Exploration Rover (MER)/Mars Science Laboratory
phenylmethylsiloxane [1]. This material has been used in many (MSL) and JUNO flight projects [3]. Within the Fermi Large Area

☆☆
The author has no financial interest in this work or any other conflicts of interest.

Distribution Statement A. Approved for public release: distribution unlimited. This material is based upon work supported by the National Aeronautics and Space
Administration under Air Force Contract No. FA8702-15-D-0001. Any opinions, findings, conclusions or recommendations expressed in this material are those of the
author(s) and do not necessarily reflect the views of the National Aeronautics and Space Administration.
E-mail address: mower@ll.mit.edu.

https://doi.org/10.1016/j.ijadhadh.2018.08.009
Accepted 7 August 2018
Available online 23 August 2018
0143-7496/ © 2018 The Author. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/BY-NC-ND/4.0/).
T.M. Mower International Journal of Adhesion and Adhesives 87 (2018) 64–72

Telescope (formerly GLAST), a photomultiplier tube was bonded to an 1.3. Scope of the present work
aluminum housing using RTV 566. The European Space Agency has also
used RTV 566 in assembly of some of its space systems—for example This study was performed to carefully characterize the thermo-
with mounting beamsplitters used for interferometers [4]. mechanical behavior of RTV 566 at reduced temperatures, because al-
though it is used extensively in aerospace applications, little informa-
tion about its behavior has been published. Thermal expansion of this
1.2. Background: crystallization in polysiloxanes material was measured over the range from +40 to −60 °C, demon-
strating the sensitivity of such measurements to the influence of multi-
Crystallization in polysiloxanes has been demonstrated through a axial constraint in such elastomeric materials. Measurements of the
few limited studies over the past sixty years, and typically occurs during elastic modulus of RTV 566 in both bending and tensile deformation
cooling but it can occur during heating, after supercooling [5]. Crys- modes were conducted over the temperature range from +40 to
tallization is an exothermic process, usually with an associated volu- −115 °C. Creep and relaxation measurements were also made, de-
metric decrease and mechanical stiffness increase. It is seldom nu- monstrating the importance of these viscoelastic behaviors particularly
cleated during brief-duration cooling (as in typical testing of at reduced temperature (−80 °C). A primary focus of this investigation
mechanical properties), but either slow cooling or prolonged exposure was to reveal whether changes in the mechanical behavior of RTV 566
to cold can trigger crystallization. evolve while it is subjected to cold temperatures for extended periods.
Though crystallization in polysiloxanes is not widely recognized To that end, measurements of the tensile modulus were performed
within the wide community that uses it in mechanical applications under isothermal conditions at temperatures as low as −85 °C, for
(even those subjected to low temperatures) several incidental ob- durations sometimes exceeding 60 h. Volumetric contraction under si-
servations of such behavior have been made. While studying gas per- milar isothermal conditions was measured, providing additional evi-
meability characteristics of silicone rubber, Zhang stated that “standard dence that this PDMS material experiences crystallization when held at
silicone rubber (VMQ) tends to crystallize at −50 °C” [6]. Rey et al. cold temperatures for long periods.
performed DSC measurements showing crystallization of RTV 141 at
−83 °C [7]. Using an interferometer to measure volumetric changes, in 2. Characterization of RTV 566 “conventional” behavior
1950 Weir, Lesser, and Wood observed behavior they ascribed to
crystallization occurring between −60 ° and −67 °C in Dow-Corning 2.1. Thermal expansion
Silastic X-6160 and in General Electric 9979G silicone rubbers, and
between −75 ° and −85 °C in Silastic 250 [8]. Their measurements 2.1.1. Methods
demonstrated that, for these materials, volume change varied from Specimens of RTV 566 material were prepared according to proce-
approximately 3 to 7% during crystallization. Electron micrographs of dures recommended by the manufacturer (Momentive Corp.). A precise
replicas of crystallites formed in PDMS at −110 °C were produced by ratio of 1 part curing agent (alkyl tin carboxylate]) to 1000 parts by
Marei et al. in 1970 [9]. More recently, direct scanning electron mi- weight resin was mixed thoroughly (using a FlackTek speedmixer and
crographs of crystals formed in PDMS at −70 °C were produced by techniques to introduce the catalyst to the resin), degassed and cast into
Sundararajan [10]. sheets with thicknesses ranging from 1 to 3 mm. These sheets of ma-
Extensive characterization of the low-temperature mechanical and terial were cured at room temperature (~23 °C) for several weeks be-
thermal behavior of a polysiloxane used in components included in fore testing.
nuclear weapons was performed by The Los Alamos National Thermal expansion of RTV 566 material was measured using a TA
Laboratory (LANL) [11]. The material (DC745U, produced by Dow Instruments Q400 thermomechanical analyzer (TMA). Note that this
Corning), consists of dimethyl, methyl-phenyl, and methyl-vinyl si- study did not explore variation in thermal expansion as a function of
loxane repeat units [12]. Chemical analysis and additional thermal/ material batches and thermal history, though such effects have been
mechanical tests were performed by Lawrence Livermore National La- observed previously. Initial thermal expansion measurements were
boratory (LLNL) [13]. Results from differential scanning calorimetry conducted in “penetration mode,” using specimens with thickness of
(DSC) tests conducted by LANL on DC745U indicate it has a glass 1.7 mm and lateral dimensions of approximately 5 ×5 mm. For these
transition temperature (Tg) of about −122 °C. During increasing tem- measurements, a contact force of 0.05 N was applied with the probe. The
perature ramps, an endothermic peak at −55 °C occurred, which is specimen temperature was held constant at −80 °C for 15 min at the
associated with the melting of crystallites. A subsequent cooling at 5 °C/ beginning of each test, and was then ramped up at 1 °C/min to 80 °C.
min revealed a crystallization temperature of about −61 °C. Mechan-
ical testing showed an increase in shear modulus by as much as two 2.1.2. Measurements
orders of magnitude, with a strong dependence on temperature. This A typical plot showing measured specimen dimension change is
particular polysiloxane experiences crystallization at approximately shown in Fig. 1a, along with the computed coefficient of thermal ex-
−50° after a dwell time of about two hours. At colder temperatures, pansion (CTE, alpha). In penetration mode, the apparent CTE of this
crystallization begins to take place much more rapidly. material at room temperature is about 400 E-6/ °C. As a consequence of
Exploration of potential crystallization in several silicones at low the very high bulk modulus of this material (nearly incompressible), the
temperatures was performed by The Aerospace Corp in 1977 [14] as expansion coefficient is influenced strongly by constraint from material
part of an investigation to determine the root cause of failure of a diode surrounding the contact area, hence measurements were also per-
crystal used in a “GPS solar array” bonded with a polysiloxane (RTV DC formed in a tensile mode to create a geometry with much less con-
6-1104). The diode crystal had routinely survived quench-cooling to straint. Specimens with approximate dimensions of
−195 °C, well below the glass transition temperature of the RTV. 0.35 × 3.5 × 13 mm were loaded in tension with 0.02 N force. The
However, when the assembly was cooled slowly (10 °C/min), the diode resulting CTE exhibited with the 50× lower geometric “constraint
crystal cracked. DSC measurements performed at this rate showed that factor” is much lower (about 220 E-6/ °C at room temperature) as seen
“substantial crystallization occurred in the vicinity of −58 °C.” Several in Fig. 1b. This result will be discussed more fully in a subsequent
other commercial polysiloxanes were included in the Aerospace study, section of this report. Note that the negative slope of the expansion
including RTV 560, 566, and 567; DSC measurements of those materials measured in penetration mode (compression) may be the result of
indicated no crystallization occurred at the cooling rates used. competing creep strain.

65
T.M. Mower International Journal of Adhesion and Adhesives 87 (2018) 64–72

Fig. 1. Thermal expansion measurements of RTV 566 in (a) penetration mode and (b) tensile mode, demonstrating effect of constraint due to highly incompressible
nature of this elastomer.

2.2. Complex modulus

2.2.1. Methods
Mechanical stiffness of RTV 566 materials was measured under a
variety of conditions, using two different TA Instruments dynamic
mechanical analyzers (DMAs): a Q800 instrument, which was used to
subject specimens to double-cantilever bending, and an RSA G2, which
was operated in tensile mode. The same lot of material used to measure
CTE was used for these initial mechanical measurements. In the canti-
lever bending tests, specimens with width of 12.5 mm, thickness of
approximately 3 mm, and length of 35 mm were deflected with an
amplitude of 200 µm at an oscillation frequency of 1 Hz. In tensile tests,
specimens with thickness of 1.8 mm, width of 3.8 mm, and length of
20 mm were subjected to strain amplitude of 0.05%, oscillating at 1 Hz.
Temperature sweeps for both types of measurements were executed at a
linear rate of 1 °C/min. Fig. 3. Effect of vacuum bakeout (two days at 65°C) on elastic modulus as
determined by DMA in cantilever bending mode.

2.2.2. Measurements
Mechanical stiffness data measured with the dynamic mechanical Fig. 3 show that vacuum bake-out appears to raise the elastic modulus
analyzers using standard temperature sweeps are plotted in Fig. 2. The by 10 to 20%.
bending results show a peak in the “tan delta” (ratio of loss modulus to
storage modulus) at about −108 °C, which is one indication of the glass 2.3. Creep and relaxation
transition temperature of polymer materials. (The peak of tan delta is
often as much as 10 °C higher than the peak of the loss modulus.) At 2.3.1. Methods
temperatures below Tg, the storage modulus (elastic, or real part of Although polysiloxane (RTV 566) can endure extremely high strains
complex modulus) rises so rapidly with decreasing temperatures that and can be considered to be a “hyperelastic” material, it is a polymer
with both instruments, the force limit was reached and the runs ter- and thus exhibits viscoelastic characteristics as well. The tensile mod-
minated. Results from the tensile measurements show peaks in the loss ulus is not a constant value when under continuous strain, but rather
modulus at about −115 °C, providing a reliable indication of the Tg. relaxes over time at characteristic rates dependent upon temperature
Influence of the effect of vacuum baking upon the elastic modulus of and, to a lesser extent, strain. Similarly, the polymer material experi-
RTV 566 was explored using cantilever bending. The data plotted in ences creep strain while under constant load. These effects could

Fig. 2. Dynamic mechanical analyzer measurements of the behavior of RTV 566 in (a) cantilever bending mode and (b) tensile deformation.

66
T.M. Mower International Journal of Adhesion and Adhesives 87 (2018) 64–72

Fig. 4. (a) Dynamic mechanical analyzer (Q800) used to measure creep behavior (b) of RTV 566 under tensile stress of 0.5 MPa, at constant temperature of −80 °C.

influence measurements of modulus conducted over an extended period extended durations to explore the possible evolution of crystallization
of time, so testing to indicate their potential effects was conducted. phenomena. A dynamic mechanical analyzer (TA Instruments RSA G2)
Sheets of RTV 566 material prepared using methods described be- was connected to a continuous supply of liquid nitrogen to make such
fore were cut into specimens with dimensions of 19.5 × 3.7 × 1.8 mm, operation possible, and provisions were made to reduce formation of
to be measured in tensile mode in a Q800 DMA, as shown in Fig. 4a. For frost and ice within the equipment. A nearly-sealed plastic bag housed
creep testing, stress of approximately 0.5 MPa was applied during iso- the entire instrument and was filled continuously with dry nitrogen gas
thermal conditions at −80 °C. Relaxation tests were performed in a supplied at 70 °C, as shown in Fig. 6.
similar manner, under isothermal conditions with temperatures ranging Each measurement began by cooling specimens at a rate of 1 °C/min
from +40 to −80 °C, while under constant tensile strain of 25%. from 40 °C to the designated cold temperature, followed by soaking for
24 to 72 h. The modulus was measured by fluctuating the tensile strain
at a frequency of 1 Hz, with a peak amplitude of 0.5%. During testing,
2.3.2. Measurements
specimens were subjected to a superimposed, constant axial strain
During creep testing at −80 °C, RTV 566 exhibited non-linear be-
ranging from 0 to 25%. Specimen gage length was 20 mm, with typical
havior immediately, with strain rising quickly and becoming linear with
specimen width of 4 mm and thickness of 1.8 mm. A video camera
time as shown in Fig. 4b. The creep strain continued to increase line-
within the environmental chamber indicated that little or no frost
arly, until after 4 h under load the strain reached a value of 19%,
formed on specimens during testing.
showing no indication of diminishment in the creep rate. Relaxation
behavior of RTV 566 is quite dependent upon temperature, as shown in
the data plotted in Fig. 5. At room temperature, the modulus drops by
3.1.2. Measurements
about 5% when subjected to a constant strain of 25% for 1 h, and
The first specimens tested in the RSA were cut from a sheet fabri-
thereafter remains nearly constant. At reduced temperatures, the tensile
cated using RTV 566 material mixed in our laboratory and cured at
modulus relaxes much more rapidly. For example, at −80 °C, modulus
room temperature for several weeks. Shown in Fig. 7 are data recorded
relaxation of over 12% occurs during the first 20 min; thereafter, the
while measuring the tensile modulus of a specimen of this material.
rate of relaxation becomes linear without reaching a plateau.
During the ramp down in temperature from +40 °C to −85 °C, the
tensile modulus increased from about 9 to 15 MPa. Note that this spe-
3. Low temperature, extended duration behavior cimen was subjected to a superimposed, constant tensile strain of 25%,
which elevates the measured modulus, relative to the behavior mea-
3.1. Tensile modulus sured while under little strain (about 5 MPa at room temperature).
While held at a constant temperature of −85 °C, an increase in the
3.1.1. Methods elastic modulus began to occur after about 1 h. After soaking for 6 h at
Measurements of the tensile modulus of RTV 566 were conducted this temperature, a fairly rapid rise in modulus occurred such that after
under isothermal conditions (“soak”) at reduced temperature for 24 h, the modulus reached 175 MPa. Thereafter, the modulus increased

Fig. 5. (a) Relaxation modulus of RTV 566, under tensile strain of 25% at constant temperature of −80 °C. (b) Comparison of stress relaxation rates as a function of
temperature.

67
T.M. Mower International Journal of Adhesion and Adhesives 87 (2018) 64–72

Fig. 6. RSA G2 enclosed in sealed bag continuously backfilled with heated, dry nitrogen gas to prevent buildup of frost during extended duration, low-temperature
testing.

Fig. 7. Elastic tensile modulus of RTV 566 while under constant superimposed Fig. 8. Elastic tensile modulus of RTV 566 material mixed and immediately
axial strain of 25% and soaked at −85 °C. Specimen material mixed and im- cured in our laboratory, under constant superimposed, axial strain at indicated
mediately cured in our laboratory. temperatures.

at a slower rate and did not reach a plateau with a steady-state value. modulus to begin to increase while under isothermal conditions at
This growth in modulus appears to be a physical manifestation of the −75 °C. Decreasing the soak temperatures to −80 and −85 °C sig-
polysiloxane material experiencing a very slow crystallization process. nificantly accelerated the onset of modulus increase and caused higher
While the test specimen was in a state exhibiting high modulus, magnitudes of modulus to be reached, so that by 48 h, over 200 MPa
distinct fluctuations in the force measured by the instrument evolved was reached for specimens soaked at −85 °C while stretched 25%.
and continued to exist while the specimen was in the changed state. Use of adhesive resins mixed with curing agents and frozen before
When observed initially, repeated tests were performed to ensure spe- curing by commercial firms is becoming increasingly prevalent. We
cimens were securely clamped in the grips. Special wave washers were therefore included in this study a batch of specimens fabricated with
installed beneath the clamp screws, to provide continued clamping RTV 566 resin that had been mixed with curing agent, frozen im-
force when the specimen thickness shrank, but fluctuations in the mediately and stored at −80 °C. The first specimen fabricated from this
measured load continued to be exhibited. A plausible explanation for material, when tested isothermally (at −80 °C), displayed such rapid
this phenomenon is that the evolving material microstructure experi- rise in elastic modulus that we tested multiple specimens from material
enced slipping between lamellar crystallites and polymer chain seg- mixed in our laboratory to confirm that instrumentation and data
ments as the crystallization process progressed. We note that to confirm processing were accurate. The resulting data shown in Fig. 9 clearly
the temporal change in measured modulus was not an artifact of the demonstrate excellent repeatability of those measurements and a pro-
cold instrument, tests were performed with no specimen installed; those nounced difference in crystallization behavior of RTV 566 material that
measurements demonstrated a “noise floor” that was four to five orders was mixed and immediately cured, versus a different batch of material
of magnitude below the forces generated while specimens were tested. that was premixed and frozen in the uncured state for approximately
Influence of temperature and applied strain were initially explored ten weeks. Further investigation of this difference in behavior is illu-
with our lab-mixed RTV 566, as indicated by the data shown in Fig. 8. strated in Fig. 10, which indicates that under similar conditions of
We observe that the initial value of the elastic tensile moduli for ma- applied strain amplitude, material that had been frozen in the uncured
terials stretched only 15% are much closer to the nominal, room-tem- state consistently experienced earlier and more significant increases in
perature value (~ 5 MPa). When soaked at −75 °C, the material showed the elastic moduli than did material that was cured immediately after
no modulus increase (crystallization) within 36 h while subjected to mixing. We note here that other differences may exist between mate-
15% axial strain. Increasing the strain to 25% did cause the measured rials mixed and cured in our lab and materials generated using

68
T.M. Mower International Journal of Adhesion and Adhesives 87 (2018) 64–72

Fig. 9. Isothermal modulus comparison between premixed, frozen RTV 566 and Fig. 12. Shrinkage of RTV 566 measured in TMA under isothermal condition at
four specimens prepared from resin mixed and cured immediately. −80 °C.

3.2. Shrinkage associated with crystallization

Because the occurrence of volumetric shrinkage during crystal-


lization of polymers has been well established (for example, by Gent in
natural rubber [15] and polybutadiene [16], and by Treloar in raw
rubber [17]), the possibility of shrinkage occurring during crystal-
lization of RTV 566 was explored using our TMA. To inhibit formation
of frost within the instrument, it was enclosed in a nearly sealed bag,
continuously filled with dry nitrogen gas. Thin specimens of RTV ma-
terial (length ~ 12 mm) were subjected to tension with a very low force
(0.01 N, producing stress ~ 11 kPa) to reduce the influence of creep on
the measured strain. Temperature was ramped down at 1 °C/min to
−80 °C, where it was held constant for extended periods.
An example of measured linear dimension change is shown in
Fig. 12. These data indicate the initial thermal contraction of the ma-
Fig. 10. Measured tensile moduli of RTV 566 at soak temperature of −80 °C,
under differing levels of applied axial strain. Comparison between two different
terial, followed by a dwell period with no change in dimension. After
batches of material, one fabricated from pre-mixed, frozen resin and one from 15 h at −80 °C, the material commenced contracting at a rate which
resin mixed and cured immediately. declined with time. Isothermal shrinkage culminated after 50 h, when
this specimen demonstrated shrinkage of about 0.35%, following the
shrinkage associated with the initial thermal contraction. Isothermal
commercially mixed, frozen resin. Exploration of such possible batch-
shrinkage (associated with crystallization) measured similarly with
to-batch differences is beyond the scope of this study.
other specimens ranged between 0.35 and 0.40%.
Several measurements were conducted to demonstrate a melting ef-
fect, wherein the warming material might reach a temperature that
caused crystallites to melt suddenly, revealed by a rapid drop in the 4. Discussion
modulus. Shown in Fig. 11 are two plots from tests that applied slow
(0.5 °C/min) increases in temperature after exposure to cold tempera- 4.1. Behavior indicated by conventional testing
tures for long periods that caused elevation of tensile modulus (crystal-
lization). In each case, the modulus decreased slowly with increasing 4.1.1. Thermal expansion
temperature, and no sharp transition was observed. When the tempera- When thermal expansion of RTV 566 is measured in penetration
ture reached approximately −58 °C, the modulus returned very near to mode (Fig. 1a) with specimens having lateral dimensions much greater
the value exhibited by the materials prior to their crystallization. than the probe diameter, apparently very high values of CTE are

Fig. 11. Heating of crystallized RTV 566 material at 0.5 °C/min, after crystallization at −85 °C (left) and −80 °C (right), revealing no indication of sharp melting
transition, with return to initial modulus (melting completed) by −58 °C.

69
T.M. Mower International Journal of Adhesion and Adhesives 87 (2018) 64–72

generated. This results from the nearly incompressible behavior of the amplitude, which is consistent with the conformational change required
material creating lateral constraint upon the volume of material be- to organize crystallites into lamellar groups. With increased externally
neath the probe, resulting in expanding material being forced to move applied strains, some straightening of random, coiled polymer chain
upwards— contributing more to the linear thermal expansion measured segments occurs, helping to overcome resistance to restructuring
in the vertical direction. The extent to which this behavior occurs with caused by the existence of molecular crosslinking. Influence of me-
specimens having finite dimensions is characterized by the “shape chanical strains upon morphology of polymer crystallization has been
factor,” which is the ratio of the area of unloaded faces of components revealed comprehensively through X-ray diffraction studies [21–23]
to their loaded areas. Influence of the shape factor upon the effective that demonstrate crystallites show a preferred orientation in the di-
material properties of rubber materials has been analyzed and in- rection of the largest strains, and the amount of crystallization increases
vestigated quite thoroughly [18,19]. rapidly with increased extension [17].
Our measurements (Fig. 1b) with thin strips of RTV 566 in tensile Lower isothermal (soak) temperatures resulted in earlier onset and
mode exhibit a linear CTE of about 220 µ/°C, much lower than the ef- greater rates of modulus increase, as well as higher apparent asymptotic
fective CTE when measured in penetration mode with surrounding values of modulus being attained. The data in Fig. 8 show no increase in
material creating constraint. Note that confined compression mea- tensile modulus when RTV 566 was soaked at −75 °C while under
surements we made recently at room temperature reveal a bulk mod- strain of 15%. Increasing the applied stretching to 25% resulted in a
ulus of approximately 75 MPa during initial, trivial loading and rise to very slow rise in modulus, suggesting that crystallization can be sti-
over 1 GPa when under high (7 MPa) hydrostatic pressure. Invoking the mulated at −75 °C with application of sufficient strain. At lower tem-
idealized equations of linear elasticity, calculated values of the Poisson peratures, modulus increase (crystallization) transpired much more
ratio range from about 0.493 under light loading, to 0.4994 when under readily. Data presented in Fig. 8 show that crystallization occurs at
high pressure. Though these values do indicate that RTV 566 is, in fact, −80 °C with just 5% applied strain, developing much more rapidly and
quite nearly incompressible, it should be realized that the elastic rela- to a greater extent with higher applied strains. At −85 °C, the process of
tions used are strictly appropriate only for small strains, and further- crystallization in RTV 566 begins soon after reaching isothermal con-
more, break down when the Poisson ratio approaches a value of 0.50. ditions. While at −85 °C and 25% strain, the tensile modulus reached
210 MPa after 48 h. Note that this temperature is approximately 30 °C
4.1.2. Elastic modulus above the glass transition temperature of RTV 566.
The data presented in Fig. 3 indicate RTV 566 has a room-tem- Unlike glass transition phenomena, crystallization in polymers is a
perature flexural modulus of about 4 MPa. With decreasing tempera- very time-dependent process. Rates of crystallization in polymers have
ture, the modulus increases linearly until about −60 °C, where the been well characterized, finding that, in general, “as the temperature is
modulus is approximately 8 MPa. Those data also reveal a small (20%) lowered, the rate increases, goes through a maximum, and then de-
increase in modulus resulting from exposures to vacuum bakeout, a creases as the mobility of the molecules decreases and crystallization
procedure used frequently to reduce offgassing of components destined becomes diffusion controlled [21].” Bukhina and Kurlyand state that
for service in space environments. the maximum rate of crystallization in polydimethylsiloxane (PDMS) is
As the temperature of RTV 566 is reduced further, the glass tran- thought to occur at a temperature of −80 °C, “but a high rate of crys-
sition region is approached, which involves an increase in the elastic tallization in this temperature range hampers the exact determination
modulus by more than two and a half orders of magnitude. Although of (the maximum rate temperature) so the available literature data are
some interpretations of Tg utilize the peak in the tan delta curve (which not unequivocal [24].”
is about −106 °C in the data shown in Fig. 2), a more widely used While the Tg of PDMS is reported as −126 °C (e.g. [5,14]), our
approach is to identify the peak of the loss modulus as Tg, which from modulus measurements indicate Tg of RTV 566 is about 11 °C higher.
these measurements appears to be about −114 °C. (Note that peaks in This is consistent with assertions that “the presence of phenyl sub-
tan delta often are as much as 15 °C above the peak loss modulus.) The stituents (in the polymer chain) results in an increase of Tg [24]” and
modulus continues to increase with decreasing temperature, causing the only description of the chemistry provided by the manufacturer of
the instrument to terminate the measurements shown in Fig. 2 when the RTV 566 (“silanol-stopped, phenyl-containing, silicone polymer”). Ad-
force limit was reached. dition of a variety of functional groups into PDMS can inhibit its crys-
tallization due to the disruption of symmetry along the polymer back-
4.2. Low-temperature isothermal behavior bone, but substituents with phenyl have been most effective in this
regard [25]. Although the presence of phenyl groups in RTV 566 delays
4.2.1. Modulus increase its ability to undergo crystallization, causing the lack of its observation
Measurements of RTV 566 in tensile mode (Figs. 7–10) under iso- in earlier studies [14], our data demonstrate that the presence of phenyl
thermal conditions at reduced temperatures present a time-dependent in RTV 566 does not prevent its crystallization when exposed to low
growth of the elastic modulus that is consistent with the progression of enough temperatures for extended durations.
crystallization in polymers, which has historically been monitored most The melting temperature of crystallites formed in RTV 566 was in-
frequently with measurement of density change [17] or volumetric vestigated by continuing to measure the tensile behavior while slowly
shrinkage [15,16]. In all of our testing at temperatures of −80 °C and heating material from a crystallized state. The data shown in Fig. 11
below, an initial static period was followed by a rise in the elastic indicate a continuous, slow process of modulus reduction while heating,
modulus that accelerated until the fastest rate of modulus increase was with melting apparently completed upon reaching −58 °C. This gradual
established, producing a linear rise in stiffness with time. The rates of progression could be due either to discrete, sequential melting of in-
modulus increase then slowed steadily, indicating that an asymptote dividual crystallites or a systemic, global softening of all crystallites
would be reached at times greater than testing constraints allowed. This simultaneously. While the latter process would have been enabled by
process parallels the description provided by Billmeyer, explaining that the uniform sample temperature created by the slow warming condi-
“….crystallization in polymers involves the steps of (primary) nucleation tions, it is possible that, just like the crystallization process, melting is a
and relatively rapid spherulite growth, followed by a slow, kinetically diffi- localized phenomenon involving individual crystallites. In contrast,
cult improvement in crystal perfection [20].” Verzino and Fillers observed distinct melting in several crystallized
The length of the initial dwell period during isothermal conditions, polysiloxanes at narrowly defined temperatures, using DSC. More re-
where no increase in modulus occurred, was quite sensitive to both cently, melting of crystallized PDMS (silanol-terminated, with Tg
temperature and applied, superimposed constant strain. Onset of ~ −127 °C) was shown by Lund et al. (also using DSC) to occur across a
modulus increase occurred more rapidly with increasing strain broad temperature range of −55 to −35 °C [26]. Melting across such

70
T.M. Mower International Journal of Adhesion and Adhesives 87 (2018) 64–72

wide temperature range is similar to what we have demonstrated with under these conditions reached a stable value of 0.35% after 50 h. The
RTV 566, which is also “silanol stopped.” time-dependency of the isothermal contraction is closely analogous to
The question may arise whether the mechanical properties of RTV change in modulus measured under similar conditions (low strain at
566 that has experienced crystallization become restored after re- −80 °C), including: the dwell period wherein little change occurs, the
turning to the amorphous state. To address this question, measurement varying rate of property change, and the time required for a stable state
of the tensile modulus as a function of temperature was performed on to develop.
two specimens after they had been crystallized extensively at −85 °C. Gent reported volumetric shrinkage during crystallization of natural
In both cases, the resulting data closely match the measurements of rubber and polybutadiene rubber to be about 2.3% and 3.4%, respec-
virgin material, as plotted in Fig. 2. Thus, it appears that the amorphous tively. Volumetric contraction of raw rubber, reported by Treloar,
polymer regains its original physical properties after being temporarily reached a maximum of approximately 2% after 300 h at 0 °C while
transformed to a crystalline state. being stretched 25% No published data for shrinkage of PDMS asso-
ciated with crystallization is seemingly available. The lower value of
4.2.2. Possible effects of freezing uncured resin shrinkage reported here may be simply attributed to the different
Tensile modulus measurements plotted in Figs. 9, 10 indicate that chemical composition and configuration of PDMS, as well as the pre-
RTV 566 material that is premixed and frozen in the uncured state, then sence of a small fraction of iron oxide filler2 in RTV 566.
cured after an extended period of time has a greater capacity to undergo
crystallization than does material that is mixed and immediately cured. 5. Conclusions
A plausible explanation for this behavior is that storing the mixed, but
uncured resin at cold temperature (−80 °C) prior to curing somehow Despite widespread use in many optical instruments of the poly-
reduces the catalyst's ability to promote crosslinking of the polymer siloxane, elastomeric adhesive material marketed under the tradename
chains. Lower crosslink density would permit conformational changes RTV 566, very little data concerning its thermomechanical behavior
to occur more readily because polymer chains would more easily uncoil have been published in open literature. Reported here are some results
and slip past each other, with fewer side groups bonded. This mobility from extensive characterization of the mechanical behavior of RTV 566
enables the folding of chains into lamella (often strain assisted) that performed at temperatures ranging from ambient to −120 °C. In ad-
provide the necessary precursors for formation of crystallites during the dition to conventional testing using temperature sweeps, the present
crystallization process [27]. investigation sought evidence that RTV 566 experiences a change in
Potential degradation of the efficacy of condensation curing agents structure, from an amorphous state to a crystalline nature, when ex-
caused by cold storage of mixed but uncured silicone resins appears to posed to cold temperatures for prolonged periods.
be a topic that has not been addressed in the open literature. It is, Measurements with a dynamic mechanical analyzer demonstrated
however, common practice to keep reactive components separate so that RTV 566 has an elastic flexural modulus of approximately 4 MPa at
that partial crosslinking does not occur. Although low-temperature room temperature. The modulus rises linearly with reducing tempera-
storage will certainly retard reaction rates, it is likely that some che- ture, nearly doubling as the temperature reaches −60 °C. As tempera-
mical reaction is still occurring at temperatures near −80 °C. Any lo- ture is reduced below −85 °C, the modulus rises very rapidly as the
calized, partial crosslinking may consume some nearby reactive end- glass transition region is encountered, with Tg shown to be about
groups and thus form isolated, non-reactive nodules. These nodules can −114 °C. Tensile relaxation measurements with a DMA demonstrated
later disrupt the normal crosslinking process when the resin is returned the viscoelastic behavior of RTV 566, with the rate of relaxation in-
to room temperature. creasing with reductions in temperature. This is consistent with the rise
Additional differences may exist between the RTV 566 materials in the loss modulus (and tan delta) at lower temperatures. Thermal
produced with resin mixed and frozen by a vendor and materials mixed expansion measurements in a TMA illustrated the effect that volumetric
and immediately cured in our own lab. Possible batch-to-batch varia- constraint can have on nearly incompressible elastomeric materials:
tions may have existed in the chemical composition of the different slender specimens in tension exhibited a CTE of approximately 220 µ/
resin sources, as well as the possible inhibition of curing mechanisms °C, but when short and wide specimens were tested in penetration
caused by freezing mixed but uncured resin. It is beyond the scope of mode, the apparent CTE value was much higher.
this study to perform comprehensive studies to explore the effects of Data produced by this study present manifestations of crystal-
batch-to-batch variations upon thermomechanical properties of RTV lization occurring in RTV 566 under isothermal conditions at reduced
566 in either the amorphous or partially crystallized states. temperatures for long periods of time. Some principal supporting evi-
dence includes these observations:
4.2.3. Shrinkage
Monitoring of crystallization in polymers has historically been • During isothermal measurements at temperatures of −80 °C and
performed indirectly using measurements of density change or heat below, after a quiescent dwell period, the elastic modulus increased
capacity. Direct evidence of formation of spherullites has been obtained with accelerating rates until nearly asymptotic levels were attained.
with polysiloxanes having chemical composition and molecular struc- This process is consistent with published descriptions of crystal-
tures that enable crystallization at warm temperatures (from ambient to lization in polymers, involving nucleation and rapid growth of
over 100 °C) using optical [28] or electron microscopy [29]. Char- spherulites followed by culmination of the changes.
acterization of strain-induced crystallization in natural rubber has been • Increase in the tensile elastic modulus can be as great as a factor of
performed widely using X-ray diffraction [23], but at ambient tem- 40 (from about 5 MPa to over 200 MPa) when the material is ex-
peratures or above. Since the aforementioned analytical techniques are posed to −85 °C for 48 h, while subjected to a constant, super-
not feasible at low temperatures for the extended periods needed to imposed axial strain of 25%. This is not a glass transition phenom-
promote crystallization in RTV 566, we sought further evidence of this enon, as this testing temperature is approximately 30 °C above the
behavior by measuring material shrinkage. Tg for this material.
Change in length of RTV 566 specimens was measured with the • Isothermal increase in the elastic modulus of RTV 566 is sensitive to
TMA at a temperature of −80 °C,1 while under an applied tensile stress the level of applied mechanical strain, which is consistent with
of about 11 kPa. The data shown in Fig. 12 indicate that shrinkage molecular orientation providing favorable nucleation sites that

1 2
This is the low-temperature limit of the instrument. Added by the manufacturer.

71
T.M. Mower International Journal of Adhesion and Adhesives 87 (2018) 64–72

accelerate the initiation and growth process of crystallization. models for rubbers, Jun, Spain, 2013; 511-514.

• Linear shrinkage in RTV 566 was observed during isothermal con- [8] Weir CE, Leser WH, Wood LA. Crystallization and second-order transitions in sili-
cone rubbers, NBS Research Paper RP2084 (44), U. S. Department of Commerce;
ditions at −80 °C in a TMA; with multiple tests, shrinkage of ap- 1950.
proximately 0.35% was measured. [9] Marei AI, et al. The strength and supermolecular structure of siloxane elastomers.
Polym Sci U S S R 1969;9(11):2162–70.
[10] Sundararajan PR. Crystalline morphology of poly(dimethysiloxane). Polymer
Acknowledgements 2002;43:1691–3.
[11] Ortiz-Acosta D. Low temperature studies on DC745U pressure pads, Report LA-UR-
This material is based upon work supported by the National 13-25885. Los Alomos, NM: Los Alamos National Laboratory; 2013.
[12] Ortiz-Acosta D. Historical material analysis of DC745U pressure pads, Report LA-
Aeronautics and Space Administration under Air Force Contract No. UR-12-23612. Los Alomos, NM: Los Alamos National Laboratory; 2012.
FA8702-15-D-0001. Any opinions, findings, conclusions or re- [13] Maxwell RS, et al. Baseline and lifetime assessments for DC745U elastomeric
commendations expressed in this material are those of the author(s) components, Report UCRL-TR208929. CA: Lawrence Livermore National
Laboratory; 2005.
and do not necessarily reflect the views of the National Aeronautics and
[14] Verzino WJ, Fillers RW. Low-temperature differential scanning calorimetry of
Space Administration. Discussions with Dr. Ethan Parsons and John polysiloxanes, Report SAM SO-TR-78-107. El Segundo, CA: The Aerospace
Wellman of MIT Lincoln Lab, concerning behavior of elastomers, are Corporation; 1978.
much appreciated. [15] Gent AN. Crystallization in natural rubber. IV. Temperature dependence. J Polym
Sci 1955(XVIJI):321–4.
[16] Gent AN, Zhang L-Q. Zhang, Strain-Induced crystallization and strength of elasto-
Data availability mers. I. cis-1,4-Polybutadiene. J Polym Sci: Part B: Polym Phys 2001;39:811–7.
[17] Treloar LRG. The physics of rubber elasticity. 3rd ed. 22. USA: Oxford University
Press; 1975.
The raw/processed data required to reproduce these findings can [18] Kelly JM, Lai J-W. The use of tests on high-shape-factor bearings to estimate the
not be shared at this time due to sponsor limitations and security re- bulk modulus of natural rubber. Seism Isol Prot Syst 2011;2(1):21–33.
strictions imposed by the U.S. government. [19] Horton JM, Tupholme GE, Gover MJC. Axial loading of bonded rubber blocks. Trans
ASME 2002;69:836–43.
[20] Billmeyer FW. Textbook of polymer science. 3rd ed. 283. NY: Wiley; 1984.
References [21] Tosaka M. Strain-induced crystallization of crosslinked natural rubber as revealed
by X-ray diffraction using synchrotron radiation. Polym J 2007;39:1207–20.
[22] Huneau B. Strain-induced crystallization of natural rubber: a review of X-ray dif-
[1] Salama AM, Rowe WM, Yasui RK. Stress analysis and design of silicon solar cell
fraction investigations. Rubber Chem Technol 2011;84(3):425–52.
arrays and related material properties, NASA Technical Report TR-32-1552.
[23] Nie Y, et al. Features of strain-induced crystallization of natural rubber revealed by
Pasadena, CA: Jet Propulsion Laboratory; 1972.
experiments and simulations. Polym J 2017;49:309–17.
[2] Yoder PR. Mounting optics in optical instruments. 2nd ed. 359. Bellingham, WA:
[24] Bukhina MF, Kurlyand SK. Low-temperature behavior of elastomers. Netherlands:
SPIE; 2008.
VSP; 2007.
[3] Ramesham R. Reliability and qualification of hardware to enhance the mission as-
[25] Zlatanic A, et al. Suppression of crystallization in polydimethylsiloxanes and chain
surance of JPL/NASA projects, In: Proceedings of the 40th international conference
branching in their phenyl-containing copolymers. Macromolecules
on environmental systems, AIAA; 2010.
2017;50:3532–43.
[4] Va Veggel AA, Nijmeijer H. Stable mounting of beamsplitters for an interferometer.
[26] Lund R, et al. Dynamical and structural aspects of the cold crystallization of poly
Precis Eng 2009;33:7–17.
(dimethylsiloxane) (PDMS). Macromolecules 2008;41:1364–76.
[5] Dupont A. Characterization of silicones, in Silicones in industrial applications
[27] Nitta KH. On the orientation-induced crystallization of polymers. Polymers
(Chapter 2). In: JaegerGleriaeditors. Inorganic polymers. NY: Nova Science; 2007.
2016;8(6):229–41.
[6] Zhang H. The permeability characteristics of silicone rubber, 2006. In: Proceedings
[28] Magill JH. Crystallization of Poly (tetramethyl-p-silphenylene)-siloxane (TMPS)
of SAMPE Fall technical conference, “Global advances in materials and process
polymers, Part 2. J Polym Sci 1967;5:89–99.
engineering”, coatings and sealants section, November 6 – 9, Dallas, TX; 2006.
[29] Haller MN, Magill JH. Morphology of polysiloxanes crystallized from the melt. J
[7] Rey T, et al. Effects of temperature on the mechanical behavior of filled and unfilled
Appl Phys 1969;40:4262–5.
silicone rubbers. In: Proceedings of the 8th European conference on constitutive

72

You might also like