Sunteți pe pagina 1din 9

European Polymer Journal 49 (2013) 2214–2222

Contents lists available at SciVerse ScienceDirect

European Polymer Journal


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

Where is the glass transition temperature


of poly(tetrafluoroethylene)? A new approach by dynamic
rheometry and mechanical tests
Gérard Calleja a, Alex Jourdan b, Bruno Ameduri a, Jean-Pierre Habas a,⇑
a
Institut Charles Gerhardt, Equipe «Ingénierie et Architectures Macromoléculaires», UMR CNRS 5253, CC 1702, Université de Montpellier 2,
ENSCM, Place Eugene Bataillon, 34095 Montpellier, France
b
AREVA – CH/DRD, BP 44-26701 Pierrelatte Cedex, France

a r t i c l e i n f o a b s t r a c t

Article history: Polytetrafluoroethylene (PTFE) has been used for many years in different application fields
Received 27 December 2012 due to its outstanding chemical and physical properties. But, the value of its glass transition
Received in revised form 19 March 2013 temperature is still today a matter of controversy and very different values are proposed in
Accepted 30 April 2013
the literature. This paper proposes to answer to this scientific question using dynamic
Available online 11 May 2013
mechanical measurements. First, the viscoelastic properties of PTFE are described on a
large temperature range and the influence of the shearing frequency is carefully investi-
Keywords:
gated. Then, the effects produced by the polymer annealing on its thermomechanical
Annealing
Dynamic mechanical measurements
behavior are detailed. This study comforts the idea that PTFE amorphous phase should
Glass transition temperature be considered as comprised of two distinct regions. The first one named ‘‘mobile amor-
Mobile amorphous fraction phous fraction’’ (MAF) is able to relax at low temperature (T = 103 °C). The other one is
PTFE specific of the macromolecular segments present at the boundaries between crystalline
Relaxation and amorphous domains. Due to the close vicinity of the crystallites, these macromolecular
segments present a more restricted mobility. The corresponding phase is designated as the
‘‘rigid amorphous fraction’’ (RAF) and its mechanical relaxation produces itself at higher
temperature (T = 116 °C). Actually, this latter value is strongly dependent on the material
crystallinity degree. In particular, it is shifted to higher temperature after occurrence of a
recrystallization that is accompanied by a further reduction of the RAF’s dynamic. Instead,
the characteristics of the MAF relaxation are poorly affected. Tensile tests also support that
the ‘‘real’’ Tg of the polymer is located at low temperature. All these results have been com-
pared to those of the literature to propose a real scientific discussion and to understand the
origin of somewhat contradictory interpretations.
Ó 2013 Elsevier Ltd. All rights reserved.

1. Introduction of the refractive index, permittivity, dissipation factor and


water absorption, as well as excellent weather durability
Fluorinated polymers are high value-added materials and resistance to oxidation. Hence, they can find relevant
for various applications, due to their unique properties applications in many fields of high technology such as
such as the thermal stability, the chemical inertness (to or- aeronautics, microelectronics, engineering, chemical
ganic solvents, oils, water, acids and bases), the low values industry, optics, textile finishing, automotive industry,
houseware, chemical processing, medical devices, architec-
tural fabrics, and wiring insulation [1–4]. Among these
⇑ Corresponding author. Tel.: +33 4 67 14 37 80; fax: +33 4 67 14 40 28.
polymers, polytetrafluoroethylene, PTFE, is by far, the most
E-mail addresses: gerard.calleja@gmail.com (G. Calleja), alex.jourdan@
areva.com (A. Jourdan), bruno.ameduri@enscm.fr (B. Ameduri),
produced and used macromolecule endowed with excep-
jean-pierre.habas@univ-montp2.fr (J.-P. Habas). tional properties (low friction coefficient, poor wear

0014-3057/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.eurpolymj.2013.04.028
G. Calleja et al. / European Polymer Journal 49 (2013) 2214–2222 2215

resistance and abrasion resistance) [5]. PTFE has been pro- currently observed during the glass transition of a poly-
duced by different companies such as Asahi, Daikin, Du- mer. This author also reported the same conclusions after
Pont, Dyneon, or Solvay Specialty Polymers under the conducting stress relaxation tests [13]. Unfortunately, in
FluonÒPTFE, PolyflonÒ, TeflonÒ, HostaflonÒ, and AlgoflonÒ both studies, the experiments were only carried out at
tradenames, respectively [3–5], to name a few. temperatures above room temperature. In other words,
Since its discovery by Plunkett in 1938, PTFE has been the domain where the c process occurred itself was not ex-
the topic of a comprehensive literature supported by the plored. After achieving the conduction of plasticization
necessity to determine the origin and the limits of its pecu- experiments on PTFE, Starkweather [14] also suggested
liar behavior. If the semi-crystalline character of the poly- the assignment of a relaxation as a glass transition. In par-
mer is universally accepted, the exact value of the glass ticular, he observed that such a relaxation was able to shift
transition temperature (Tg) is still today subject to contro- with the volume fraction of diluent while the position of c
versies. Indeed, the reported values of PTFE’s Tg range from relaxation was almost constant. Extending his researches
110 °C to 130 °C with intermediate values such as 70 or to the characterization of different semicrystalline fluoro-
50 °C [5,6]. The difficulty to assess this critical tempera- polymers by thermally stimulated currents, this author
ture by calorimetry could be considered as a first element attributed the c process of PTFE to localized and coopera-
responsible of this open debate [7]. At the same time, other tive motions of a few CF2 units. Consequently, the occur-
techniques currently used with polymers led to results that rence of this relaxation in the ac dielectric response of
can be difficult to interpret. An illustration of this complex- various poly(TFE-co-HFP) random copolymers was conse-
ity was encountered with viscoelastic measurements usu- quently judged logical [15]. While analyzing the relaxation
ally convenient to evaluate the Tg of a polymer. In 1959, behavior of PTFE in the temperature domain including the
McCrum [8] registered the rheological behavior of PTFE a process, Wortmann [16] also considered this latter pro-
versus the temperature. The thermomechanical profile cess as relevant of a glassy/amorphous transition.
showed the presence of four main relaxations that ranged First divergence in the location of PTFE glass transition
between 200 °C and 380 °C. Characterizing the rheologi- appeared with researches conducted by Durrell et al.
cal response of samples with distinct crystallinity degrees, [17]. These authors suggested to model the Tg evolution
this author showed that two transitions were related to the of a TFE copolymers series versus their chemical composi-
crystalline fraction of the polymer. The first one located at tion using the Fox equation:
ca. 330 °C (Tm), was easily assigned to the polymer’s melt-
1=T g ¼ w1 =T g1 þ w2 =T g2 ð1Þ
ing point and that appeared consistent with the sharp drop
of the mechanical rigidity. The other one, observed in the where w1 and w2 stand for the weight fraction of each
20–30 °C range, was related to the reorganization of PTFE’s monomer incorporated, while Tg1 and Tg2 are assigned to
crystalline structure. The two last rheological transitions the glass transitions of the corresponding homopolymers.
observed in the PTFE thermomechanical profile were qual- A good agreement between the experimental and calcu-
ified as second order transitions. As their respective inten- lated values was obtained considering that the Tg of PTFE
sities were found to increase when the crystallinity was close to 50 °C, i.e. an intermediate position between
decreased, these transitions were assigned to the response Tc and Ta. Later, Boyer [18] reported similar results. In dif-
of amorphous domains [9]. But, the comparable ampli- ferent studies devoted to the rheological characterization
tudes of their corresponding relaxation peaks did not allow of perfluoropolyethers, Marchionni et al. [19,20] also sug-
the authors to precise the exact nature of the mechanisms gested to consider PTFE’s glass transition temperature
involved in these transitions. Prudently, the relaxation ob- close to 75 °C.
served at 110 °C (Tc) was attributed to small sections of Another research trend associated the low temperature
the macromolecule whereas that registered at 130 °C (Ta) c relaxation (110 °C) to the glass transition temperature
was assigned to large molecular segments. Later, Eby and of PTFE. Different experimental approaches were under-
Sinnot [10] named these relaxations as Glass I and Glass taken using mechanical, acoustical or thermodynamic
II transitions, respectively, but the authors did not explain techniques [21–23]. More recently, Rae and Dattlebaum
the presence of two glass transitions in PTFE. [24] investigated the properties of PTFE in compression
In 1963, Tobolsky et al. [11] suggested to consider the a on a wide temperature zone and for different crystallinity
transition centered at 110 °C as the PTFE glass transition. degrees. Low-crystalline samples showed a little difference
The authors came to this conclusion after conducting the in their mechanical responses with strain-rate or tempera-
stress relaxation experiments at temperatures higher than ture above 100 °C. A reverse situation was observed be-
the polymer’s melting point. Indeed, the best description of low this critical temperature interpreted as being close to
the results was achieved from the Williams–Landel–Ferry the polymeric Tg. Fossati et al. indirectly led to the same
equation in which the Tg of the polymer was close to conclusion in a study devoted to the sorption and perme-
110 °C. Indeed, the description of the experimental data ation of hydrocarbons in a poly(TFE-co-perfluoromethyl-
with this semi-empirical relation was judged satisfactory vinylether) copolymer, MFA [25]. Using PTFE as a
considering the value of PTFE close to 110 °C. Using an material reference for their modeling, they used the com-
experimental approach based on the measurement of the mon assumption that the polymer’s Tg was above room
linear thermal expansion coefficient at different tempera- temperature. But, surprising results motivated the authors
tures, Araki [12] also proposed to set PTFE’s Tg at 123 °C to examine the reverse situation. Under this latter assump-
(396 K) due to the discontinuity in the evolution of this tion, a satisfactory representation of sorption data was
parameter at this temperature that resembled that obtained.
2216 G. Calleja et al. / European Polymer Journal 49 (2013) 2214–2222

Table 1 summarizes the different techniques used to as- 2.2. Rheological characterization
sess the glass transition temperature of PTFE. The uncer-
tainty about the exact value of the PTFE’s Tg may appear The different viscoelastic experiments were performed
surprising considering that this material received much using a stress-controlled dynamic rheometer (AR2000Ex
attention by the scientific community [26]. But, most stud- model from TA). This apparatus was equipped with an
ies were devoted to the description of its exceptional prop- environmental testing chamber to allow the registering
erties in different fields related to physics and chemistry as of the complex shear modulus G ¼ G0 þ j G00 under precise
mentioned above. Basic information such as the exact nat- control of the temperature. The component G0 , usually
ure of the molecular mechanisms at the origin of both c called ‘‘storage modulus’’, represents the mechanical rigid-
and a relaxations were regarded with less attention in ity of the sample (i.e. its elastic contribution). The loss
spite of their usefulness. Thus, it was of interest to revisit modulus G00 relates to the dissipated mechanical energy
‘‘basic’’ concept such as the glass transition of this polymer due to molecular motions in the material. The behavior
by using a novel approach involving dynamic mechanical of the PTFE versus temperature was investigated from
and aging tests. First, to go further in the description of 150 °C up to the molten state using rectangular torsion
the rheological behavior of PTFE, the thermomechanical re- geometry. The different thermomechanical tests were car-
sponse of this material was registered on a large tempera- ried out at a heating rate of 3 °C min1 and at a fixed oscil-
ture zone and for different values of the shearing lating angular frequency x but ranging from 1 to
frequency. Then, the effects produced by a thermal treat- 100 rad s1 to examine the influence of x on the tempera-
ment (annealing) on the polymer dynamic mechanical ture position of the different relaxation peaks. The repro-
properties were investigated. Finally, a real discussion ducibility of the rheological results was checked by
was suggested by comparison of our results with the data repeating twice the analyses.
already published in separate articles to propose reliable
structure–properties relationships. 2.3. Mechanical characterization

2. Experimental The tensile properties of PTFE were assessed using a


universal material test machine from Zwick (model Z010)
2.1. Material equipped with a 5 kN load cell. The tests were performed
at ambient temperature at a constant speed of 5 mm/min
PTFE samples were kindly provided by Solvay Specialty according to ISO527 norm.
Polymers and marketed under the AlgoflonÒ trademark.
This neat polymer was received under a rectangular sheet 2.4. Physical treatment
that was re-cut under rectangular tablets using automated
saw working at low velocity. Typical dimensions of the To complete the scientific understanding of the rheo-
specimens were 40 mm  8 mm  1 mm. This geometry logical relaxations, some PTFE samples were exposed at
was judged well suited for the viscoelastic characterization the temperature of 105 °C in dry atmosphere for one
of the polymer in solid state. The same grade was also re- month. The annealing effects on the polymer’s thermome-
ceived under dog-bone form for tensile tests (width of chanical behavior were investigated using the same exper-
10 mm and thickness of 4 mm). imental conditions as previously described.

Table 1
3. Results and discussion
Different approaches used to assess the Tg of PTFE with the attached
corresponding values.
3.1. Thermomechanical analyses of initial PTFE
Technique Tg Author
value
(°C) First rheological experiments were conducted on PTFE
‘‘as received’’ i.e. without any physical or chemical treat-
Dynamic mechanical analysis 110 McCrum [8]
(DMAs) +130 Starkweather [14]
ment. Fig. 1 shows the evolution of the viscoelastic proper-
Wortmann [16] ties of the polymer as a function of temperature and at the
Steady or transient rheometry +110 Tobolsky [11] constant angular frequency x = 1 rad/s. The evolution of
Araki [13] the storage modulus G0 = f(T) agrees with that reported in
Mechanical measurements 110 Woodward and Sauer
the literature by McCrum [8] or Hintzer and Löhr [27].
[21]
Rae and Dattelbaum [24] Moreover, it describes a more extended temperature zone
Dilatometry (CTE) +123 Araki [12] than that detailed in Starkweather’s work [14]. Undeni-
Thermally stimulated currents +130 Sauer [15] ably, the thermomechanical profile can be qualified as
(TSCs) being complex since four transitions are shown on the rhe-
Calorimetry and 50 Durrell [17]
thermodynamics 110 Lau [23]
ological curves.
Calculation 50 Boyer [18] Two transitions are well known and their interpretation
75 Marchionni [19,20] does not suffer from any form of controversy. That ob-
Acoustic 110 Kvacheva and served from 50 °C up to 32 °C is named b and currently
Perepechko [22]
related to the crystalline transitions that produce them-
Permeability measurement <20 Fossati [25]
selves in PTFE [25]. Indeed, X-ray experiments have
G. Calleja et al. / European Polymer Journal 49 (2013) 2214–2222 2217

1.E+10

1.E+09

Moduli G' & G" (Pa)


1.E+08

1.E+07

1.E+06

1.E+05
-150 -125 -100 -75 -50 -25 0 25 50 75 100 125 150 175 200 225 250 275 300 325
Temperature (°C)

Fig. 1. Thermomechanical analysis of PTFE with G0 : d and G00 : 4 (x = 1 rad/s).

already shown in the past that such a thermoplastic poly- result is of first importance because it reveals that none of
mer can exhibit three solid phases at atmospheric pressure these relaxations can be confused with ‘‘simple’’ rheological
with two first-order crystal–crystal transitions that occur secondary relaxations. Indeed, associated to local segmental
at 19 °C and 30 °C [24,28]. The first one is known to be motions in the polymer chain, secondary relaxations induce
characteristic of the transition from a highly ordered tri- mechanical losses that appear as rheological peaks on the
clinic structure to a hexagonal crystal. More precisely, curve characteristic of the dissipative energy i.e. the loss
the macromolecule slightly untwists since it passes from modulus G00 . Due to its reduced amplitude, this phenome-
a 13/6 helical conformation (13 CF2 groups are equally non has a poor effect on the G0 modulus in contrast to the
spaced in 6 turns) to a 15/7 helix. The second crystalline rheological evidence of the glass transition temperature
transition corresponds to the evolution to a pseudohexag- [29–34]. Indeed, in this latter process, the molecular dynam-
onal crystal due to the further untwisting and that is ics is usually characterized by cooperative motions of higher
accompanied by a rise in the crystalline cell from 13 to molecular segments which affect the polymer mechanical
15 carbon atoms. In rheology, the differentiation of these rigidity [35,36]. Then, at this stage of that present discus-
crystalline transitions is more difficult because they pro- sion, PTFE really seems to exhibit two glass transition do-
duce themselves in neighboring temperature zones. Never- mains as suggested by Eby and Sinnot [10]. But, the
theless, the width and the asymmetry of the relaxation existence of both transitions remains unexplained.
peak in the temperature range associated to these crystal- Then, to investigate further the characteristics of c and
line changes seem to be consistent with the presence of a processes, another series of thermomechanical analyses
two transitions. The former one seems to prepare itself were performed on new PTFE samples in the solid state
well before the critical temperature of 19 °C since the in- to analyze more deeply the influence of the value of the
crease of the G00 curve from 50 °C reveals a progressive shearing angular frequency on the position of each relaxa-
growing in the local molecular mobility. The sudden drop tion. As shown in Fig. 3a and b, both c and a mechanical
between 20 °C and 30 °C of the storage modulus curve peaks occur at higher temperatures when x increased.
agrees with the reported crystalline transition from tri- The evolution of the temperatures taken at the maxi-
clinic to hexagonal phase. Another universally accepted mum of each relaxation peak can be described using the
transition is reported above 320 °C where the values of model proposed initially by Glasstone et al. [55] to detail
both moduli suddenly decrease. This phenomenon is char- the effect of the temperature on the diffusion phenomena:
acteristic of the polymer melting as shown by the presence
of an endothermic peak in PTFE calorimetric analysis [28].
x ¼ A expðEa =RTÞ ð2Þ
Both c and a relaxation peaks mentioned in the intro- where R stands for the gas constant (8.32 J/mol/K), T (K)
duction can also be observed on the G00 curve [14]. They ap- the absolute temperature at which the maximum of G00 is
pear in a clearer form by plotting the evolution of the loss observed for the corresponding angular frequency x
factor d (defined by tan d = G00 /G0 ) versus the temperature (rad/s), Ea the activation energy (kJ/mol) and A the preex-
(Fig. 2) [8]. Named also ‘‘internal friction’’, this loss factor ponential factor (rad/s). In other words, two apparent acti-
is really influenced by the nature and amplitude of molec- vation energies can be calculated from the plot of log x
ular motions [27]. The maxima of each relaxation peak are versus the inverse of the temperature taken at the maxi-
respectively taken down at Tc = 103 °C and Ta = 116 °C. mum of each corresponding peak (Fig. 4).
It is interesting to note that both relaxations also induce a This result is currently observed for secondary transi-
significant decrease of the storage modulus G0 (Fig. 1). This tions and this may be contradictory with previous remarks.
2218 G. Calleja et al. / European Polymer Journal 49 (2013) 2214–2222

0.18 β

0.16

0.14
α
0.12
γ
0.10
tan δ

0.08

0.06

0.04

0.02

0.00
-150 -125 -100 -75 -50 -25 0 25 50 75 100 125 150 175 200 225 250 275 300
Temperature (°C)

Fig. 2. Evolution of the tan d of PTFE as a function of the temperature.

equation [37]. Actually, due to the limited evolution of


(a) 0.19 w = 100 rad/s
the angular frequency value, it was not possible to discrim-
0.17 w = 32 rad/s
w = 10 rad/s
inate which kind of law the most appropriate. The exploi-
0.15 tation of the results according Arrhenius dependence gives
w = 3.2 rad/s
0.13 w = 1 rad/s the respective activation energy: Ec = 105 kJ/mol and
0.11
Ea = 444 kJ/mol. These latter values are in the same order
tan δ

of magnitude as that proposed for the characterization of


0.09
PTFE from dielectric spectroscopy [15,38]. Besides this the-
0.07 oretical treatment, it is interesting to note that the change
0.05 in frequency did not induce any shoulder formation or
0.03 peak dissociation. This means that c and a processes are
governed by specific major molecular mechanisms.
0.01
-150 -125 -100 -75 -50
Temperature (°C)
3.2. Characterization of PTFE after annealing at T = 105 °C

(b) 0.18 Fig. 5 depicts the evolution of the dynamic mechanical


0.16 properties, namely tan d, of a PTFE sample after annealing
at the temperature of 105 °C for 72 h. This temperature
0.14 was fixed at a value comprised in the temperature domain
0.12 associated to the a relaxation. To make possible a better
tan δ

understanding of the changes induced by the thermal


0.10 treatment, the thermomechanical response of the ‘‘initial’’
w = 100 rad/s polymer is also drawn.
0.08 w = 32 rad/s
w = 10 rad/s While the annealing step produced little effects on the c
w = 3.2 rad/s relaxation, it induced more significant changes on both po-
0.06
w = 1 rad/s
sition and shape of the other rheological peaks. First, the b
0.04 peak that is characteristic of different crystalline transi-
50 75 100 125 150 175
tions in the PTFE undeniably exhibits higher amplitude.
Temperature (°C)
Nevertheless, the temperature associated to the maximum
Fig. 3. Influence of the shearing angular frequency x on the position of c
is unaffected. Second, the a relaxation peak is shifted to
and a peaks measured on tan d = f(T) and represented by (a) and (b) series, higher temperature after annealing: Ta value increased
respectively. from 116 °C up to 124 °C. In addition, at the same time,
the melting enthalpy assessed by DSC also increased from
31 J/g to 37 J/g.
Indeed, one of c or a relaxations is suspected to be associ- The higher enthalpy of fusion seems to reveal that the
ated with the glass transition of the polymer. Then, in this annealing step has induced an increase of the polymer
case, the corresponding temperature (Tc or Ta) would crystallinity. This hypothesis is coherent considering also
more likely obey the William–Landel–Ferry (WLF) the growing intensity of the transition b. In fact, the
G. Calleja et al. / European Polymer Journal 49 (2013) 2214–2222 2219

5.5

4.5

3.5
ln ω = -53.378 / Tα + 137.36
ln ω (rad/s)
2.5

ln ω = -12.663 / Tγ + 74.042
1.5

0.5

-0.5
2.4 2.6 2.8 3 3.2 3.4 3.6 3.8 4 4.2 4.4 4.6 4.8 5 5.2 5.4 5.6 5.8
1000/T (K-1)

Fig. 4. Evidence for the Arrhenius dependence of Tc and Ta temperatures versus the shear angular frequency x.

0.22

0.20

0.18

0.16

0.14

0.12
tan δ

0.10

0.08

0.06

0.04

0.02

0.00
-150 -125 -100 -75 -50 -25 0 25 50 75 100 125 150 175 200
Temperature (°C)

Fig. 5. Non-isothermal tan d curves of the initial PTFE (solid line) and after annealing at 105 °C for 3 days (dotted line).

phenomenon described here points out the ‘‘crystal perfec- domains located in the close vicinity of the crystallites
tion’’ already observed in different semi-crystalline poly- [45]. In other words, the a process is only characteristic
mers [39–42]. This mechanism seems perfectly valid with of the mechanical relaxation of these constrained amor-
PTFE. Indeed, from X-ray diffraction technique, Jain and phous areas and the higher Ta rise is the direct conse-
Vijayan [43] investigated the effects produced by a thermal quence of the inner stresses increase due to the
aging conducted at T = 150 °C on PTFE structure and they development of crystalline zones.
showed an increase of the polymer’s degree of crystallinity. Our hypothesis seems consistent with results of various
Similar results were also obtained by Yamamoto and Hara surveys in which many semicrystalline polymers should be
[44] for an exposure temperature close to Ta. considered as three-phase systems [46–50]. In this ap-
The understanding of the evolution of both c and a proach initially used to explain the discrepancies between
peaks must be highlighted now. A reduction of their ampli- the degrees of crystallinity obtained from different tech-
tude was expected as reported by McCrum [8] who as- niques (calorimetry, X-ray diffraction, NMR or Raman
signed these relaxations to polymer amorphous phase. spectroscopy), the amorphous phase is modeled as made
But, in this present work, the main noticeable change is of two distinct domains. The first one is named ‘‘mobile
the increase of Ta. A possible mechanism is to imagine that amorphous fraction’’ (MAF) and corresponds to the classi-
the expansion of the crystalline phase is accompanied by a cal representation of macromolecular chains randomly ar-
reduction of the molecular mobility of the amorphous ranged in space. The second one is defined as ‘‘rigid
2220 G. Calleja et al. / European Polymer Journal 49 (2013) 2214–2222

amorphous fraction’’ (RAF) and is associated with the Tg must be taken at the value characteristic of the MAF,
interphase between the crystalline and the mobile amor- i.e. at Tc = 103 °C.
phous phase. The existence of this intermediate interfacial Dividing PTFE according to three phases also agrees
region is justified by the continuation across the phase with the researches conducted by Dlubek et al. [51]. These
boundaries of the macromolecules (Fig. 6). Due to its par- authors investigated the local free volume of PTFE and its
ticular position, the RAF is characterized by a molecular copolymer based on TFE and perfluoro(propyl vinyl ether),
mobility that is intermediate between that of MAF and PFA, on a large temperature scale. Comparing the results of
crystalline domains. It can be reasonably believed that this pressure–volume–temperature experiments with data ob-
description is applicable to the case of PTFE represented tained by positron annihilation lifetime spectroscopy, they
schematically in Fig. 6: if the a relaxation is assigned to led to the conclusion that a fraction of the amorphous
the RAF, the c relaxation is likely to be characteristic of phase was restricted in its segmental mobility due to the
the MAF. Considering that the main glass transition tem- incorporation of polymer chains into the crystals. Indeed,
perature of the polymer is defined as being characteristic they proposed to split the amorphous phase into two dif-
of the amorphous phase of the ‘‘perfect’’ glass, the PTFE’s ferent subdomains with specific properties according to
an approach that resembles the MAF and RAF concepts de-
scribed above. After exploitation of their PVT experiments,
they evaluated the volumetric glass transition temperature
of PTFE lower than 85 °C. However, these authors did not
mention any volume transition in the vicinity of 130 °C.
This first result seems consistent with our own assignment
of the c relaxation as being the rheological evidence of
PTFE’s glass transition. Another information raised from
Dlubek’s work points out that the constrained amorphous
phase which is formed during the polymer crystallization
from the melt is characterized in the first times by a smal-
ler specific free volume compared to the MAF. But, the re-
stricted segmental mobility in the RAF limits the
contraction of the corresponding domains and finally the
RAF presents a larger free volume than that specific of
the MAF. To our opinion, this finding should be analyzed
considering Starkweather’s work [14] that reports the
effects of PTFE exposure versus different fluids (hexafluo-
ropropylene dimer, perfluorodimethylcyclohexane, chloro-
form, etc.). In this latter research, the a peak assessed by
DMA was the only one affected by the chemical treatment.
Its shift to lower temperature region was interpreted as
(1) (2) (3) significant of the polymer plasticizing due to the fluid dif-
Fig. 6. Schematic illustration of PTFE structure. (1): mobile amorphous fusion in the polymer matrix. Consequently, Starkweather
phase; (2): rigid amorphous phase; (3): crystalline phase. proposed to interpret the a relaxation as a characteristic of

40

35

30
Stress (MPa)

25

20

15

10

0
0 25 50 75 100 125 150 175 200 225 250 275 300 325 350 375 400 425
Strain %

Fig. 7. Tensile behavior of PTFE measured at 25 °C with the displacement velocity of 5 mm/min.
G. Calleja et al. / European Polymer Journal 49 (2013) 2214–2222 2221

the polymer’s glass transition. But, the same evolution can on localized molecular motions. Some researchers already
also be considered as a pure consequence of the higher free proposed to relate both c and a relaxations to the amor-
volume of the RAF compared to the MAF. Indeed, the liquid phous phase of PTFE considering that their respective
is likely to diffuse in an easier way in the constrained amplitudes were reduced by an increase of the polymer
amorphous phase plasticizing the corresponding macro- crystallinity favored by low cooling rates from the molten
molecular segments. state. However, in our study, the annealing of the polymer
Other key elements seem to support our interpretation at T = 105 °C made it possible a better discrimination of
of both c and a rheological relaxations. The tensile behav- these relaxations. Indeed, this heat treatment was assessed
ior of PTFE can be regarded as an additional response. Fig. 7 as responsible of the occurrence of a cold crystallization in
illustrates that PTFE presents a ductile character at room the polymer matrix. Moreover, this latter phenomenon in-
temperature with a breaking strain that exceeds 400%. This duced the shift of the a peak to higher temperatures
feature that is responsible of the inadequate use of PTFE in whereas the c relaxation remained unaffected. This differ-
structural functions seems to be consistent with the entiation led us to consider that the amorphous phase of
assignment and centering of the polymer’s glass transition PTFE should be divided into two sub-categories. The first
at Tc. Indeed, under this hypothesis, the classical amor- one was defined as specific of the rigid amorphous fraction
phous macromolecular segments are likely to present a (RAF) located at the vicinity of the crystallites. The second
mobility high enough to induce the polymer’s ductility at one, named MAF, was described as being composed of the
ambient temperature. remaining amorphous segments characterized by a higher
Another interesting point to take into account is the molecular mobility. Consequently, the a relaxation was as-
polymer Tg value evaluated using the semi-empirical signed to the rigid amorphous phase while the c process
method proposed by Van Krevelen [52]. According to this observed at lower temperatures (110 °C) was attributed
approach, the chemical structure of the polymer can be to the mobile amorphous fraction. Different information
split into elementary chemical groups characterized by were precised to support our interpretation. First, the dis-
specific contribution Yi and molar mass Mi. Then, the poly- placement of the a peak observed after annealing was ex-
mer’s Tg is calculated from the ratio of the summation of plained as being representative of the further reduction of
the elementary Yi by the sum of the individual weight Mi the macromolecular mobility at the immediate vicinity of
according the following equation: the crystallites due to the cold crystallization. Second, the
P high mechanical ductility of PTFE at ambient temperature
Yi seemed another consistent key element to locate the poly-
Tg ¼ P i ð3Þ
i Mi mer’s glass transition temperature in the region where the
c process was observed. Third, the concept of dividing PTFE
The value of YCF2 obtained from different fluorine-derivat-
according to three phases was also found in the literature
ed polymers was evaluated to 10,500 K g/mol whereas
that reports the characterization of this polymer by PVT
MCF2 worths 50 g/mol. The final calculation led to PTFE’s
measurements. Finally, the same approach was judged
Tg close to 210 K i.e. 63 °C. Although this value does not
helpful to explain already published results about the ef-
correspond exactly to that assessed for the Tc, it seems to
fects produced by PTFE exposure versus fluorinated fluids
support the idea for which the PTFE’s glass transition is lo-
on its thermomechanical behavior.
cated at very low temperature. Actually, it is important to
keep in mind that the Tg is a practical parameter to define
the temperature domain where the glass transition is cen- Acknowledgements
tered. Indeed, this latter domain can be 50 K wide due to
the continuous contribution of the different macromolecu- The authors are grateful to AREVA NC for financial sup-
lar segments [53,54]. port and to Solvay Specialty Polymers for supplying PTFE
as a free sample.

4. Conclusion
References
This paper aims at bringing a new insight in the defini- [1] Ameduri B, Boutevin B. Well-architectured fluoropolymers:
tion and location of PTFE’s glass transition temperature. To synthesis, properties and applications. Amsterdam: Elsevier; 2004.
answer to this scientific question, we have chosen to use [2] Feiring AE. Organofluorine chemistry: principles and commercial
applications, vol. 15. New York: Plenum Press; 1994. p. 339.
dynamic mechanical measurements to study and interpret [3] Scheirs J. Modern fluoropolymers. New York: John Wiley & Sons Ltd.;
the nature of the different molecular mechanisms at the 1997.
origin of the different thermomechanical relaxations. As [4] Hougham G, Cassidy PE, Johns K, Davidson T. Fluoropolymers 2:
properties. New York: Kluwer/Plenum; 1999.
the b peak is universally accepted as being representative
[5] Ebnesajjad S. Non-melt processible fluoroplastics. Norwich
of crystalline transitions in the polymer, most of our work (NY): Plastics Design Library; 2000.
has been devoted to the investigation of c and a relaxations [6] Dobkowski Z, Zieleka M. Polimery 1999;44(3):222–5.
[7] Furakawa GT, McCoskey RE, King GJ. J Res Natl Bur Stand
that have been found controversial in the literature. Our
1952;49:273–80.
first results underlined that both rheological processes [8] McCrum NG. J Polym Sci 1959;34:355–69.
were accompanied by an important decrease of the poly- [9] Gray RW, McCrum NG. J Polym Sci 1969;7:1329–55.
mer rigidity. Actually, they were identified as characteristic [10] Eby RK, Sinnott KM. J Appl Phys 1961;32:1765–71.
[11] Tobolsky AV, Takahashi M, Katz D. J Polym Sci Part A 1963;1:483–9.
of segmental molecular motions of large amplitude. No [12] Araki YJ. Appl Polym Sci 1965;9(2):421–7.
confusion was possible with secondary relaxations based [13] Araki YJ. Appl Polym Sci 1965;9(4):1515–24.
2222 G. Calleja et al. / European Polymer Journal 49 (2013) 2214–2222

[14] Starkweather HW. Macromolecules 1984;17:1178–80. [36] Aharoni SM. J Appl Polym Sci 1976;20(10):2863–9.
[15] Sauer BB, Avakian P, Starkweather HW. J Polym Sci Pol Phys [37] Ferry JD. Viscoelastic properties of polymers. 3rd ed. New
1996;34:517–26. York: Wiley & Sons; 1980.
[16] Wortmann FJ. Polymer 1996;37(12):2471–6. [38] Starkweather HW, Avakian P, Matheson RR, Fontanella JJ, Wintersgill
[17] Durrell WS, Stump EC, Schuman PD. J Polym Sci Pol Lett MC. Macromolecules 1991;24:3853–6.
1965;3:831–40. [39] Badrinarayanan P, Dowdy KB, Kessler MR. Polymer
[18] Boyer RF. Plast Polym 1973;41:71–7. 2010;51(20):4611–8.
[19] Marchionni G, Ajroldi G, Righetti MC, Pezzin G. Polym Compos [40] Shieh YT, Lin YS, Twu YK, Tsai HB, Lin RH. J Appl Polym Sci
1991;32(3):71–3. 2010;116(3):1334–41.
[20] Marchionni G, Ajroldi G, Righetti MC, Pezzin G. Macromolecules [41] Chen SH, Wu YH, Su CH, Jeng U, Hsieh CC, Su AC, et al.
1993;26:1751–7. Macromolecules 2007;40(15):5353–9.
[21] Woodward AE, Sauer JA. Mechanical relaxation phenomena. In: Fox [42] Martinelli A, D’Ilario L, Caminiti R. J Polym Sci Part B – Polym Phys
D, editor. Physics and chemistry in the organic state. New 2005;43(19):2725–36.
York: Interscience; 1965. p. 637–723. [43] Jain A, Vijayan K. Polym Eng Sci 2007;47(11):1724–9.
[22] Kvacheva LA, Perepechko I. Sov Phys Acoust 1973;18(3):343–6. [44] Yamamoto T, Hara T. Polymer 1982;23:521–8.
[23] Lau SF, Suzuki H, Wunderlich B. J Polym Sci Pol Phys [45] Karagiannidis PG, Stergiou AC, Karayannidis GP. Eur Polym J
1984;22:379–405. 2008;44:1475–86.
[24] Rae PJ, Dattelbaum BM. Polymer 2004;45:7615–25. [46] Alsleben M, Schick C. Thermochim Acta 1994;238:203–27.
[25] Fossati P, Sanguineti A, De Angelis MG, Baschetti MG, Doghieri F, [47] Schick C, Dobbertin J, Potter M, Dehne H, Hensel A, Wurm A, et al. J
Sarti GC. J Polym Sci Pol Phys 2007;45:1637–52. Therm Anal 1997;49(1):499–511.
[26] Gianneti E. Polym Int 2001;50:10–26. [48] Strobl G. Eur Phys J 2000;E3:165–83.
[27] Hintzer K, Löhr G. Modern fluoropolymers. New York: John Wiley & [49] Schick C, Wurm A, Mohamed A. Coll Polym Sci 2001;279(8):800–6.
Sons Ltd.; 1997 [chapter 11]. [50] Wunderlich B. Prog Polym Sci 2003;28(3):383–450.
[28] Lehnert RJ, Hendra PJ, Everall N, Clayden NJ. Polymer [51] Dlubek G, Sen Gupta A, Pionteck J, Hassler R, Krause-Rehberg R,
1997;38(7):1521–35. Kaspar H, et al. Polymer 2005;46:6075–89.
[29] Pater RH. Polym Compos 1991;12(2):133–6. [52] Van Krevelen DW. Properties of polymers: their correlation with
[30] Habas JP, Peyrelasse J, Grenier-Loustalot MF. High Perf Polym chemical structure; their numerical estimation and prediction from
1996;8:515–32. additive group contributions. 4th ed. Amsterdam: Elsevier; 2009.
[31] Buchenau U. J Non-Cryst Sol 2007;353:3812–9. [53] Flory PJ. Principles of polymer chemistry. Cornell & London: Cornell
[32] Blachot JF, Chazeau L, Cavaille JY. Polymer 2002;43(3):881–9. University Press; 1953.
[33] Xiao CD, Yee AF. Macromolecules 1992;25:6800–9. [54] Kalogeras IM, Lobland HEH. J Mater Ed 2012;34:69–94.
[34] Nassiet V, Habas JP, Hassoune-Rhabbour B, Baziard Y, Petit JA. J Appl [55] Glasstone S, Laidler KJ, Eyring H. The theory of rate processes: the
Polym Sci 2006;5:679–90. kinetics of chemical reactions, viscosity, diffusion and
[35] Gibbs JH, DiMarzio EA. J Chem Phys 1958;28(3):373–83. electrochemical phenomena. New York: McGraw-Hill; 1941.

S-ar putea să vă placă și