You are on page 1of 14

Feature Article

Segmental Mobility in the Non-crystalline


Regions of Semicrystalline Polymers and its
Implications on Melting

Sanjay Rastogi,* Yefeng Yao, Dirk R. Lippits, Günther W. H. Höhne,


Robert Graf, Hans Wolfgang Spiess, Piet J. Lemstra

Detailed knowledge on chain mobility in polymers is of fundamental interest in order to


understand their mechanical properties. As a specific example, the melting behavior of semi-
crystalline polyethylene can be studied by thermal analysis and NMR spectroscopy. In ultra high
molecular weight polyethylene (UHMW-PE)
crystallised via different routes, i.e., directly
during polymerisation, from solution, or from
the melt, and melted under different protocols,
different melting processes involving detach-
ment of stems from the crystals and cluster
melting can be distinguished. Melting by the
consecutive detachment of chain stems from
the crystal substrate ultimately results in a melt
state where chain dynamics for entanglement
formation are much more restricted.

Introduction
S. Rastogi
Department of Materials, Loughborough University, Leicester- In semicrystalline polymers, crystalline regions are linked
shire LE11 3TU, United Kingdom by non-crystalline domains. Depending on the crystal-
E-mail: s.rastogi@lboro.ac.uk lisation conditions, chain stiffness, and molecular weight, a
S. Rastogi, D. R. Lippits, G. W. H. Höhne, P. J. Lemstra chain may traverse between crystals or fold back within the
Department of Chemical Engineering and Chemistry/ The Dutch parent lamella, thus influencing the number of entangle-
Polymer Institute, Eindhoven University of Technology, P.O. Box ments residing in the non-crystalline regions and their
513; 5600MB Eindhoven, The Netherlands
topology. This, in turn, has pronounced effects on chain
E-mail: s.rastogi@tue.nl
organisation, dynamics and melting, considered here for
S. Rastogi, Y. Yao, R. Graf, H. W. Spiess
Max-Planck-Institute for Polymers, Ackermannweg 10, 55128
a well studied flexible polymer - linear ultra high
Mainz, Germany molecular weight polyethylene of molar mass greater
Y. Yao than 106 g  mol1 (UHMW-PE).[1,2,3]
Department of Physics, East China Normal University, North The chain topology can be also controlled by crystallizing
Zhongshan Road 3663, 200062 Shanghai, P. R. China the polymer from solution. Below a critical overlap

Macromol. Rapid Commun. 2009, 30, 826–839


826 ß 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim DOI: 10.1002/marc.200900025
Segmental Mobility in the Non-crystalline Regions of . . .

concentration, the entanglements are reduced to an extent Tm  131 8C, i.e., 5 8C lower than the experimentally
that the solution cast films can be drawn more than observed melting point of 136 8C. Furthermore, the high
150 times in the solid state.[4] Recently, we demonstrated melting temperature of 136 8C is lost on second heating
that careful synthesis with a single site catalyst produces PE where a melting temperature of 131 8C is measured, which
where the nascent crystals are highly disentangled. Similar now coincides with the prediction of the Gibbs-Thomson
to solution cast films, the films obtained on compression of equation.
the nascent crystals, below the melting temperature, can be The melting aspects involved in nascent, melt- and
drawn in the solid state more than 150 times.[3] Distinct solution-crystallised polymers cannot be explained by
morphological differences between solution crystallised and established thermodynamic concepts alone. Therefore, in
nascent crystallised ‘‘disentangled’’ crystals exist. While in this article we feature an extended investigation into the
the solution crystallised sample lamellae of 12.5 nm are melting behavior of these polymers. We correlate the time,
regularly stacked on top of each other that allows crystal as well as the temperature, required for melting with the
thickening,[5] whereas in the nascent sample, the regular chain topology in the non-crystalline regions. The intro-
stacking of lamellae is absent. In both cases, the solid-state duced kinetic aspect of melting is probed by different
drawability is lost once the ‘‘disentangled’’ crystals are melt experimental techniques: differential scanning calorimetry
crystallised. This loss in drawability is attributed to the (DSC), temperature modulated differential scanning calori-
increase in entanglements in the non-crystalline region of metry (TM-DSC), 1H and 13C NMR spectroscopy and
the semicrystalline polymer influencing the chain topology relaxometry. The key observation in our study is that, at
between the crystalline domains. Synthesis with the highly sufficiently low heating rates, disentangled nascent poly-
active heterogeneous Ziegler-Natta catalyst on the other ethylene melts can be prepared with unusual properties.
hand, yields nascent (entangled) UHMW-PE, which is less The lowest temperature where melting is observed
drawable in the solid state (7 times), indicating the influence corresponds to the value predicted by the Gibbs-Thomson
of the synthesis conditions on the chain topology in the non- equation. This indicates that topological constraints in the
crystalline region of the semicrystalline polymer. non-crystalline domains influence the melting of the
The melting temperatures of samples crystallised during crystalline regions in the semicrystalline polymer, an
polymerisation (nascent) or from a melt of the same observation which should hold for melting of polymer
polymer are distinctly different. For example, while the crystals in general.
nascent UHMW-PE, on heating at 10 8C  min1, melts
around 141 8C, close to the reported equilibrium melting Experimental Part
temperature for polyethylene of 141.5 8C, crystals obtained Materials
via the solution route melt already at approximately 136 8C
and the melt crystallised sample at approximately 135 8C. To correlate the melting behavior of UHMW-PE to the chain
The high melting temperature for the nascent polymers has topology, four different samples were studied as specified in
Table 1. The two nascent grades differ in synthesis conditions and
been a subject of debate. Using electron microscopy and
catalyst type. The nascent entangled sample is a commercial grade
DSC, Engelen[6] et al. conclusively showed that the nascent
of Montell (1900CM) synthesised with a heterogeneous Ziegler-
crystals contained folded chains and displayed a lamellar
Natta catalyst. The nascent disentangled grade is synthesised at
thickness in the region of 12 nm. Thus, the high melting temperatures below the dissolution temperature of PE using a
temperature for nascent polymers was attributed to homogeneous metallocene catalyst.
successive thickening prior to melting. However, no The solution crystallised samples were obtained with the
experimental evidence of successive thickening was entangled nascent UHMW-PE on crystallisation from solution (for
provided. On the contrary, Kurelec et al. showed that, even example 1 wt.-% of UHMW-PE in decalin). Crystals thus obtained
on annealing close to the melting point for several hours, are disentangled and have been investigated in detail before.[5] The
the lamellar thickness of these nascent crystals does not melt crystallised samples were obtained by cooling the entangled
exceed 26 nm.[7,8] The melting temperature predicted melt at 10 K  min1.
Tensile testing was used to investigate drawability of the
from the Gibbs-Thomson equation for polyethylene[9]
materials in the solid state below the melting point. In the
(Tm ¼ 414.2–259.7/l) for a lamellar thickness l of 26 nm is
following, we denote drawable materials (>150 times) as
131 8C.[10] Furthermore, the high melting temperature of
‘‘disentangled’’ and un-drawable (<7) materials as ‘‘entangled’’.
141 8C is lost on second heating, where a melting
temperature of only 135 8C was found.[10] A similar
Experimental Techniques
discrepancy was observed between the first and second
heating run of solution crystallised UHMW-PE, where the
Differential Scanning Calorimetry (DSC)
lamellae double their initial thickness upon annealing
below the melting temperature to a maximum of 25 nm.[5] DSC was performed using a standard Perkin-Elmer DSC-7. Samples
From the Gibbs-Thomson equation we again expect of 0.7–2.0 mg mass were weighed with a precision balance and

Macromol. Rapid Commun. 2009, 30, 826–839


ß 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.mrc-journal.de 827
S. Rastogi, Y. Yao, G. W. H. Höhne, R. Graf, H. W. Spiess, P. J. Lemstra

Table 1. Molecular properties and synthesis temperature conditions of the UHMW-PE samples.

Mw Mw =Mn Melting point Draw ratio


on first heating at 120 -C
g  mol1

Nascent, Entangled 4.5  106 8 141 8C 7


6
Nascent, Disentangled 3.6  10 3.0 141 8C >150
Solution-crystallised 4.5  106 8 136 8C >150
6
Melt-crystallised 4.5  10 8 135 8C 7

encapsulated in standard (crimped) aluminium pans of known achieved by the rotor encoded REDOR (ReREDOR) technique.[13] For
mass. An identical empty pan was used as a reference. Nitrogen was the ReREDOR experiments, a commercial 2.5 mm MAS double-
purged at a rate of 25 ml  min1. The DSC was calibrated using resonance probe was used at a spinning frequency of 25 kHz. The
indium and tin. same probe was used to study the chain diffusion between non-
crystalline and crystalline regions based on a 13C exchange type
Tensile Testing experiment[12,14] under MAS at 6 kHz. The 908 pulse length in
For draw ratio determination, dumb bell-shaped samples with a the 2.5 mm MAS probe was adjusted to 2.5 ms on both channels. The
gauge length of 10 mm and a width of 0.5 mm were used. Samples temperature of the bearing gas was varied for temperature
were drawn in the solid-state at a crosshead speed of 50 mm  min1 dependent experiments.
at 125 8C. The draw ratio was determined by measuring the
Temperature Modulated Differential Scanning
separation of the ink marks prior and after the deformation.
Calorimetry (TM-DSC)
NMR Spectroscopy TM-DSC was performed using a standard Perkin Elmer DSC-7
1
H NMR experiments were carried out without sample rotation on a apparatus modified for temperature modulation.[15] For the TM-
DSC measurements, an underlying heating rate of 0.5 K  min1 was
Bruker DMX spectrometer operating at a 1H NMR frequency of
500 MHz and equipped with a special 7 mm probehead that resists used, which was low enough to ensure good linearity and stability
temperatures above 150 8C. The transverse relaxation time T2 was during measurements. Time dependent temperature-modulated
measurements were performed in quasi-isothermal mode at
measured using a two pulse sequence 908-t-1808-t-aq with a
variable t time starting from t ¼ 2 ms. The 908 pulse length was 5 ms different temperatures (110 8C, 120 8C, 130 8C, 135 8C, 136 8C,
and the repetition time was 3 s, which proved long enough for 137 8C and 138 8C) at a frequency of 12.5 mHz. The temperature
quantitative measurements. The relaxation decay was charac- amplitude TA of 53 mK was kept low enough to ensure a linear
response. Before evaluation, the empty pan run (taking the same
terised by 60 data points at properly selected echo times. This pulse
sequence was chosen because it offered the possibility to both measuring parameters and phase position) was subtracted from
qualitatively as well as quantitatively analyse relaxation of the the curve of the sample to reduce surroundings and apparatus
influences. The measured (modulated) heat flow rate signal F(T,t)
non-crystalline and crystalline components. Temperature calibra-
tion was carried out by monitoring peak separation in the 1H NMR consists of two parts: the underlying part Fu(t) and the periodic part
spectrum of glycol and the melting-induced 1H NMR line Fper.(T,t). Gliding integration over one period provides the under-
lying part Fu(t) which, when subtracted from the total measured
narrowing of a series of compounds also employed as DSC
reference materials. 1H NMR spin-spin relaxation decays were signal, yields the periodic part Fper.(T,t). The ‘‘apparent’’ heat
obtained from the total integral of the spectra after Fourier capacity is calculated from the periodic part using a mathematical
procedure described in ref. [16].
Transformation, phase and baseline correction. The relaxation
decay was analyzed by a non-linear least-square fitting. The
relaxation times were determined by fitting the relaxation data
with a sum of two or three exponentials. From the determined
relaxation times below the melting point, appropriate t-delay Results and Discussion
times were chosen for real-time monitoring of the T2 relaxation by
12 data points. In this way it was possible to follow the changes in First Melting Point of Nascent UHMW-PE
the sample with 5 min intervals.
Figure 1 shows a standard DSC run of entangled and of
All solid state NMR[12] studies were performed on a Bruker
DSX spectrometer operating at 500 MHz 1H Larmor-frequency. disentangled nascent UHMW-PE. Independent of the
Advanced NMR recoupling techniques correlating isotropic che- synthesis route, no differences in the melting peak, onset
mical shifts observed under magic angle spinning (MAS) with and end temperatures exist, though the polydispersity and
motionally averaged 1H-13C heteronuclear dipole-dipole couplings the topological constraints (estimated by differences in the
were used to study the local chain dynamics in the samples. The drawability of the two samples in the solid state[9])
measurement of 1H-13C heteronuclear dipole-dipole couplings was generated during synthesis of the two samples are

Macromol. Rapid Commun. 2009, 30, 826–839


828 ß 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim DOI: 10.1002/marc.200900025
Segmental Mobility in the Non-crystalline Regions of . . .

and (ii) by the time dependence of the melting process. The


‘‘true’’ melting temperature is usually determined by
extrapolation to zero heating rate, yielding Tm ¼ 138.4 8C
and 138.2 8C for the disentangled and entangled samples,
respectively (Figure 2). According to the Gibbs-Thomson
equation, these melting temperatures correspond to a
crystal thickness of approximately 100 nm.[19] The experi-
mental observations, however, show clearly that the
crystals do not thicken to more than 26 nm,[7,8] even when
the sample is left to anneal for several hours at 120 8C.
Similar to the nascent UHMW-PE, the solution-crystal-
lised films of the UHMW-PE also show significantly higher
melting temperatures than expected from their crystal
thickness. For example, lamellar doubled crystals with a
thickness of 25 nm melt at 5 8C above the anticipated
melting point of 131 8C.
Figure 1. DSC curves obtained on heating the nascent entangled
Thus, the Tm of solution-crystallised and the nascent
and the nascent disentangled samples at 10 8C  min1. Indepen-
dent of the synthesis route the polymer melts at 140.5 8C, close to UHMW-PEs differ by more than 5 8C, despite similar crystal
the equilibrium melting point of the linear polyethylene. thickness. Likewise, the melting temperatures of the melt-
crystallised and the nascent state of the same polymer
differ by more than 6 8C, though the crystal thickness of the
two samples is similar. Apparently, in addition to the
different. The observed high melting point of 141 8C of the crystal thickness, other structural and dynamic parameters
nascent powders is 5–6 8C higher than for melt-crystallised influence the melting process. In order to determine
UHMW-PE.[6] On heating at rates of 20 down to 0.1 K  min1, differences in local chain organisation and dynamics on
, a non-linear relationship between the peak temperature different time and length scales, solid state NMR was
and the heating rate of entangled and disentangled nascent employed.[12]
UHMW-PE polymer is observed. However, plotting the
melting temperature versus the heating rate to power 0.2,
yields a linear relationship (Figure 2). Similar observations
Solid State NMR Studies of UHMW-PE
were reported by Toda et al.[18]. However, unlike Toda et al.,
no instrumental corrections are applied to the measure- Linear polyethylenes crystallised from solution and from
ments reported here. The shift in the observed melting the melt were investigated by solid-state NMR[12] to
temperature at non-zero heating rates is attributed to probe the segmental/medium-range chain mobility.[20,21]
‘‘superheating’’. This superheating is caused (i) by thermal Advanced NMR recoupling techniques were used to detect
inertia (the transport of heat from the heater to the sample) the motionally averaged 1H-13C heteronuclear dipole-
dipole coupling.[13] The medium-range chain diffusion
between non-crystalline and crystalline regions was
monitored via 13C exchange type experiments[12,14] under
MAS.
Figure 3(a) and 3(b) show 13C CP/MAS spectra of the
solution and melt-crystallised samples, respectively. The
peak at 33 ppm arises from the all-trans conformations in
the crystalline regions.[22] Chain segments in the non-
crystalline regions adopt different conformations leading
to a peak at 31 ppm.[22] Due to differences in the
distribution of chain conformations, the position of the
peak is slightly different in the solution- and melt-
crystallised samples. Moreover, in contrast to the melt-
crystallised sample, the peak width of the non-crystalline
region in the solution-crystallised sample is rather broad
Figure 2. Measured melting peak temperatures for the entangled and reaches all the way to the all-trans peak. This is a first
and the disentangled nascent ultra high molecular weight poly- indication of extended trans conformations and reduced
ethylenes (UHMW-PE) at different heating rates (b). dynamics in the non-crystalline region of such samples.

Macromol. Rapid Commun. 2009, 30, 826–839


ß 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.mrc-journal.de 829
S. Rastogi, Y. Yao, G. W. H. Höhne, R. Graf, H. W. Spiess, P. J. Lemstra

Figure 3. The 13C CP/MAS spectra (6 kHz MAS, 2 ms CP contact) of


PEs and ReREDOR35 sideband patterns (25 kHz MAS, 3 rotor
periods recoupling) from the non-crystalline regions, recorded Figure 4. 13C MAS exchange spectra of the solution-crystallised
at T ¼ 320 K. The 13C CP/MAS spectra of (a) solution-crystallised sample (SC-PE) and the melt-crystallised sample (MC-PE): (a)
and (b) melt-crystallised PE show two components, the non- measured at 6 kHz MAS, T ¼ 320 K with different exchange times;
crystalline signal (31 ppm, red dashed line) and the crystal peak (b) measured at 6 kHz MAS, a fixed exchange time (5 s) and with
(33 ppm, green dashed line). The ReREDOR sideband pattern of different experimental temperatures.
non-crystalline, solution-crystallised and of non-crystalline melt-
crystallised PE. The underlying blue pattern are simulations with
a dipole-dipole coupling constant of DIS/2 ¼ 9.1 kHz for the waiting time two peaks are observed in the solution-
solution-crystallised sample and 6.3 kHz for the melt-crystallised
crystallised sample. The peak positions match with those of
sample.
the non-crystalline and the crystalline regions. The
intensity of the crystalline peak increases rapidly with
In Figure 3(c) and 3(d), the 1H-13C dipole-dipole sideband the exchange time. Considering the long relaxation times
spectra of solution- and melt-crystallised samples are required for the crystalline component, the build-up of the
compared. The profile of the sidebands represents the crystal peak is attributed to the exchange of chain segments
strength of the 1H-13C dipolar coupling. The broad profile of from the non-crystalline to the crystalline region. Such
the sidebands of the non-crystalline region in the solution- exchange, though very weak, is also observed in the melt-
crystallised sample compared to that of the melt-crystal- crystallised samples at a much longer exchange time of
lised sample is indicative of a stronger residual dipole- 400 s.
dipole coupling in the former. From the sideband patterns, The exchange process in linear polyethylenes has been
the strength of the residual dipole-dipole coupling of the studied earlier by Schmidt-Rohr and Spiess[14] and more
non-crystalline component was obtained[20] as 9.1 kHz for recently on UHMW samples of different morphol-
the solution-crystallised and only 6.3 kHz for the melt- ogy.[20,21,23] These studies demonstrate that the extended
crystallised sample. The higher dipole-dipole coupling in chain conformations present in the non-crystalline region
the solution-crystallised sample is attributed to extended of the solution-crystallised sample play a conclusive role in
trans conformations and more restricted local chain the cooperative chain motion, despite the restricted local
dynamics in the non-crystalline region, for details see chain mobility. The higher disorder in the non-crystalline
ref.[20, 21]. region of the melt-crystallised sample of the same polymer
To investigate the influence of local chain organisation suppresses cooperative chain diffusion despite higher local
and mobility on the cooperative chain motion between the mobility. This has implications for our understanding of
crystalline and the non-crystalline regions (chain diffu- crystal thickening. For example, the solution-crystallised
sion), 1D 13C NMR exchange spectra[12,14] have been sample exhibits enhanced chain mobility along the crystal-
recorded. Figure 4(a) shows a series of NMR spectra recorded lographic c-axis, which ultimately leads to doubling of
at a fixed temperature of 320 K for different waiting the initial crystal thickness.[5] Contrary to this, crystals in
(exchange) times, where the waiting time is the time the melt-crystallised sample of the same material
needed for the build up of the signal after saturation of the hardly thicken. Moreover, the higher chain order in the
13
C signal. Due to the large difference in the relaxation times non-crystalline regions of solution-crystallised samples
of the crystalline (>1 500 s at 320 K) and the non-crystalline together with the high translational chain mobility
regions (0.6 s at 320 K), build up of the signal for the non- explains the observed easy mechanical deformation of
crystalline region will be much faster than the build up of the crystals in the solid state. Last but not least, melting of
the signal for the crystalline region. With increasing the crystals and chain reorganisation of the chains prior to it

Macromol. Rapid Commun. 2009, 30, 826–839


830 ß 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim DOI: 10.1002/marc.200900025
Segmental Mobility in the Non-crystalline Regions of . . .

involves chain motion on different time and length scales


identified here. Therefore, the NMR findings should be kept
in mind when discussing the melting behavior of the
different samples.

Melting Kinetics in Solution-crystallised UHMW-PE


as Probed by DSC
To get more insight into the melting mechanism, annealing
experiments at temperatures below the ‘‘true’’ melting
point of 138.4 8C (determined from Figure 2) were
performed in the range from 132 8C to 138 8C. After
annealing at the requisite temperature, the sample was
cooled to room temperature and reheated (at 10 K  min1)
to 150 8C. On re-heating, two melting peaks at about 135 8C
Figure 6. Arrhenius plot of the time constants determined from
and 141 8C are observed (Figure 5). The peak at 135 8C is the relaxation time of the annealed sample for the entangled and
ascribed to the melting of that part of the sample which was disentangled nascent UHMW-PEs (see ref. [11]). The activation
molten and re-crystallised during cooling from the energies of the three different processes are realised from the
annealing temperature and the second peak at 141 8C is three different slopes of the curves in the Figure (reproduced
from ref. [11]).
ascribed to the remaining crystal domains still in the initial
nascent state. The ratio between the areas of the two peaks
changes with the annealing time at the given annealing
temperature. As the peak area is proportional to the amount
the heat of fusion) reads:
of the respective crystals, the ratio between the areas is also
a measure for the ratio between the amounts of crystals
which are re-crystallised and those which are in the initial HðT; tÞ ¼ H0 ðTÞet=tðTÞ (1)
nascent state.
From such studies,[11] it is possible to determine the where the time constant (t) can be related to an activation
relaxation times for the different melt processes of the energy (EA) by
different samples. With a Debye (Arrhenius) type time law
for the fusion process in question, the enthalpy change (i.e., t ¼ t 0 eEA =RT (2)

Figure 6 summarizes the relaxation


times at different annealing tempera-
tures. Closed symbols represent the
relaxation times for the nascent disen-
tangled UHMW-PE sample. Open sym-
bols in the Figure represent the relaxation
times for the nascent entangled sample.
For the disentangled sample above
135.5 8C only one single relaxation time
is found, whereas below this temperature
two relaxation times are observed. Con-
trary to that, the entangled polymer
shows only one relaxation time over
the whole temperature range of 133–
138 8C. In fact, altogether three regions
with different slopes are observed indi-
Figure 5. Heating run of the disentangled UHMW-PE, after annealing at 135 8C for cating the involvement of three different
different annealing times. The annealing times are mentioned in the corresponding
activation energies in the two different
figures. The first peak at 133 8C corresponds to melting of the crystals which crystallised
from the melt obtained on annealing. The second peak at 141 8C corresponds to the temperature regions (see Table 2). The
nascent crystals. The high temperature melting peak disappears on annealing the activation energies vary from
sample at 135 8C (reproduced from ref. [11]). 5.0  1.0  103 kJ  mol1 (slope (c)) for

Macromol. Rapid Commun. 2009, 30, 826–839


ß 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.mrc-journal.de 831
S. Rastogi, Y. Yao, G. W. H. Höhne, R. Graf, H. W. Spiess, P. J. Lemstra

Table 2. Activation energy determined from the Arrhenius plot shown in Figure 6 and Figure 7.

Activation energy Number of CH2 Chain length Lamellae thickness Stems


(T103 kJ  mol1)a) units involved involved prior to melt involved

nm nm

Nascent entangled 4.3  1.0 (c) 1 600  370 202  50 260,0 6–10
2.4  0.2 (b) 900  90 110  10 260,0 4–5
0,0
Nascent disentangled 5.0  1.0 (c) 1 850  370 235  50 26 7–11
2.1  0.2 (b) 780  50 99  7 260,0 4
0,0
0.6  0.05 (a) 220  20 28  2 26 1
Solution-crystallised 4.2  1.0 (c) 1 560  370 197  50 250 6–10
0.7  0.05 (a) 250  20 32  2 250 1

a)
(a),(b),(c) refer to the three slopes determined from the Figure 6 and 7. Activation energies in the Table are determined from slopes in the
Figures.

temperatures above 136 8C via 2.1  0.2  103 kJ  mol1 substrate and their cooperative diffusion and the activation
(slope (b)) to 0.6  0.05  103 kJ  mol1 (slope (a)) for energy (c) refers to the breakdown of parts of the crystal by
temperatures below 136 8C, respectively. simultaneous randomisation of at least 7–8 stems.
The three activation energies suggests to consider three Contrary to the disentangled polymer, the entangled
different melt processes in the disentangled nascent polymer does not show slope (a), suggesting the absence of
UHMW-PE, assigned to detachment/removal of chain melting through consecutive detachment of single chain
stem(s) from the crystal, followed by the diffusion of the stems from the crystalline substrate. Apparently this
detached chain segment into the melt. From physical process is not possible in such samples due to the
chemistry it is known[24] that the molecular motion in constraints in the non-crystalline regions. However, the
liquids is an activated process and the viscosity follows an slopes of (b) and (c) of the entangled and the disentangled
Arrhenius type law. The detachment of stems from the samples are similar, indicating that melting of the
crystal can be seen as a dynamic process with activation entangled sample occurs in clusters of at least several
energy proportional to the number of CH2-groups involved. chain stems. Such a melting process further confirms the
The rate of desorption of molecules from any surface presence of topological constraints in the nascent
follows a similar law.[24] Within the frame of this model, the entangled UHMW-PE compared to the disentangled
lowest activation energy (a) can be assigned to a UHMW-PE.
cooperative detachment of only a couple of chain stems Figure 7 shows the measured time constants of the
from the crystal surface and their diffusion into the relaxation process for solution cast films at different
surrounding melt. The intermediate activation energy (b) temperatures. In the explored temperature region,
then refers to the simultaneous detachment of a larger similar to Figure 6, two distinct slopes (a) and (c) are
number of stems from the crystalline substrate and their observed. Slope (a) refers to the activation energy of
diffusion and finally the highest activation energy (c) refers 0.7  0.05  103 kJ  mol1 required for the detachment and
to the cooperative breakdown of larger parts of the crystal diffusion of one chain stem of approximately 30 nm length,
lattice because of instability in the thermodynamic sense. a value in accordance with the measured crystal thick-
The low activation energies determined from the slopes (a) ness.[8] Slope (c) refers to the activation energy of
and (b) suggest the involvement of a ‘‘new melting’’ 4.2  1.0  103 kJ  mol1, a value comparable to that of
behavior, whereas the slope (c) refers to the conventional the respective process in the nascent UHMW-PE. This high
(thermodynamic) melting at higher temperatures. activation energy is again assigned to the breakdown of the
More quantitatively, the activation energy (a) can be crystal lattice. The absence of the slope (b) in Figure 7
assigned to the detachment of a single stem 28 nm (this suggests that within the temperature region of 128 to
has been calculated considering the activation energy 130.5 8C melting in these solution cast films mainly occurs
for detachment of one CH2 group and its diffusion into the by removal of single chain stems from the crystal substrate.
melt equal to 2.7 kJ  mol1 and the CC distance in the The lowest and the highest activation energies (repre-
orthorhombic lattice along the c-axis equal to 0.127 nm).[25] sented by slopes (a) and (c)) of the disentangled nascent
The intermediate activation energy (b) then refers to the sample in Figure 6 and the solution-crystallised sample in
simultaneous detachment of 3–4 stems from the crystalline Figure 7 are similar. The crystal thicknesses of the

Macromol. Rapid Commun. 2009, 30, 826–839


832 ß 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim DOI: 10.1002/marc.200900025
Segmental Mobility in the Non-crystalline Regions of . . .

DSC method can be found in a review article of


Wunderlich.[26]
By use of TM-DSC Höhne et al. recently showed[15] that
the magnitude of the complex apparent heat capacity of the
nascent disentangled UHMW-PE in the pre-melting region
is higher than the heat capacity obtained by conventional
DSC in that region. This indicates that chain reorganisation
occurs in the solid state and contributes to the apparent
heat capacity. The magnitude of the ‘‘excess heat capacity’’
is lower on second heating suggesting that the process is
irreversible and is best observed on first heating of the
nascent sample. Quasi-isothermal TM-DSC experiments
revealed that this apparent heat capacity is time depen-
dent. Evaluation of the relaxation times of the processes
contributing to the apparent heat capacities is possible
Figure 7. Arrhenius plot for solution-crystallised UHMW-PE under the assumption that the involved relaxation
samples. The different relaxation times at the given annealing processes are of Arrhenius type (i.e., the contributions to
temperatures show two activated processes. the apparent heat capacity follow an exponential decay). It
should be noted, however, that due to the assumptions and
the large uncertainty of the determined relaxation times
the results should be used with care.
disentangled nascent and the solution-crystallised samples Quasi-isothermal TM-DSC measurements on the disen-
are comparable, and both samples can be deformed tangled nascent UHMW-PE samples are performed at
mechanically in the solid state. This suggests that the different temperatures below the melting peak and the
activation energy required for the respective processes, change of the apparent heat capacity in time is plotted in
namely the removal of single chain stems from the surface Figure 8. The respective excess heat capacity function
followed by their diffusion in the melt (slope (a)), as well as is fitted with a sum of 2 or 3 exponentials (a1et/t1 þ
the breakdown of larger parts of the crystal lattice (slope (c)), a2et/t2þ a3et/t3) and the results of the best fit are plotted
is largely independent of the differences in the crystal- in Figure 9. The approximately linear behavior in this so-
lisation conditions of the two samples. The melting called Arrhenius-plot allows determination of activation
processes are distinctly different, however, from the energies for chain reorganisation in the solid state as
melting behavior observed for the melt-crystallised sam- 30  10, 60  10 and 500  150 kJ  mol1, respectively, for
ples, where only one melting process occurs at 135 8C (at a the three processes found in this sample. Comparing the
heating rate of 10 K  min1) without any time dependence two lower activation energies (30  10, 60  10 kJ  mol1)
of melting at low heating rates or at annealing tempera- with the enthalpy change from solid to liquid of one CH2-
tures below 135 8C. To follow the processes in the crystal group (4.11 kJ  mol1[25]) the corresponding processes are
lattice prior to melting (e.g., reorganisation and pre-
melting) temperature modulated DSC (TM-DSC) measure-
ments on the entangled and disentangled nascent UHMW-
PE were performed and have been described elsewhere.[8]

Reorganisation in Solid State prior to Melting as


Probed by Temperature Modulated DSC (TM-DSC)

From the above described experiments it is evident that


different processes are involved in the melting of crystals.
Temperature-modulated differential scanning calorimetric
(TM-DSC) investigations provide the opportunity to sepa-
rately measure the individual contributions of different
processes to the ‘‘apparent’’ heat capacity. In the quasi-
isothermal operation mode it is possible to study the
Figure 8. Apparent heat capacity from four quasi-isothermal TM-
contributions of such processes dependent on time. DSC measurements in the melting region of the disentangled
Detailed information about such processes and the TM- nascent UHMW-PE (frequency 12.5 mHz, period 80s, TA ¼ 53 mK).

Macromol. Rapid Commun. 2009, 30, 826–839


ß 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.mrc-journal.de 833
S. Rastogi, Y. Yao, G. W. H. Höhne, R. Graf, H. W. Spiess, P. J. Lemstra

is annealed close to the onset melting temperature. In this


case melting proceeds by successive detachment of chain
stems from the crystal substrate and diffusion into the
neighboring melt. The minimum temperature required for
this process is in good agreement with the melting
temperature determined from the Gibbs-Thomson equa-
tion. At higher annealing temperatures (>135.5 8C) another
melting process occurs where larger clusters, exceeding the
size of one stem, are co-operatively involved (for details see
ref. [11]).
The relaxation times required for the successive detach-
ment of chain stems from the crystal substrate and the
Figure 9. Activation diagram of the three processes determined diffusion into the surrounding matrix (>150 min) are
from Figure 8. The slope gives the activation energy (solid symbol: considerably longer than the relaxation times for melting in
exothermic, open symbols: endothermic). The stars represent the clusters. Apparently, at standard heating rates, melting
time constants determined from the slope (a) of the Figure 6. mainly proceeds via the cluster melting involving 7–8 chain
This is attributed to the melting and diffusion of the stems from
stems. The experiments clearly show that such a cluster
the crystal sides. The activation energy is determined from the
slopes. melting of the nascent UHMW-PE leads always to a high
melting temperature of 141 8C, normally ascribed to
extended chain crystals. In the second heating of the
nascent samples (i.e., melt-crystallised), no time dependent
likely to involve less than twenty methylene units, a melting on annealing is observed and the melting
number lower than the units of a single stem in the crystal. temperature is 5 8C lower than the melting temperature
Thus the two processes observed in TM-DSC cannot be of the nascent polymer during the first run. When combined
attributed to real melting, but suggest local dislocations with the NMR findings on differences in the mobility of the
which enable chain diffusion related to the a-relaxation non-crystalline regions of the melt and solution/nascent-
which occurs just in this temperature region. These findings crystallised samples,[20,21] these findings suggest that the
and the low activation energy are also in agreement with above described ‘‘cluster melting’’ (slope (b) in Figure 6) is
the NMR studies of chain diffusion indicated above and absent in the melt-crystallised samples. Therefore, we
described in detail elsewhere.[20,21,23] conclude that distinction in the melting behaviour and
The third process, observed above 130 8C in the TM-DSC temperature also reflects topological differences of the
experiments (unfilled symbols along line (c), in Figure 9) is chains in the non-crystalline regions of the semicrystalline
linked to irreversible melting. It seems to be one of the polymers rather than differences in crystal size only.
melting processes seen in the conventional DSC studies The dynamics of crystal melting is usually related to the
mentioned in the section above. Indeed, the data points increasing number of defects in the crystal lattice,
obtained from slope (a) of Figure 6 inserted as star ultimately leading to complete breakdown of the lattice
symbols () in Figure 9 coincide with the TM-DSC data, at a certain concentration of defects.[27] In semicrystalline
indicating that the two relaxation processes are of similar polymers, where a chain within the lattice is connected
origin. Thus, as above we attribute the third process from to chain segments in the non-crystalline regions, the
the TM-DSC experiments to the detachment of one chain dynamics of melting requires cooperative motion of the
stem from the crystal substrate and its diffusion to the melt. chain segments both within the lattice and in the non-
crystalline region. Differences in tight or loose chain folds
connecting chains in the non-crystalline and crystalline
regions will therefore have implications in the breakdown
Origin of the High Melting Point of
of the crystal lattice. This concept is shown schematically in
Nascent and Solution-Crystallised UHMW-PE
Figure 10. In nascent UHMW-PE with adjacent or tight folds,
If we compare the low activation energies of 30 and a greater number of CH2 groups have to move
60 kJ  mol1 for the first and second process in the TM-DSC cooperatively (Figure 10(b), 10(c)), thus melting requires
experiment with the activation energy required for the the number of involved CH2 groups to be higher than
detachment of a single chain stem (600 kJ  mol1), it is those present in a single chain stem residing within the
concluded that the two processes with low activation crystal lattice. The cooperative melting will involve several
energy are associated with chain diffusion within the chain stems connected by tight folds, resulting in a higher
crystals rather than detachment of the chain. The third melting temperature than that predicted from the Gibbs-
process takes place when the disentangled polymer sample Thomson equation.

Macromol. Rapid Commun. 2009, 30, 826–839


834 ß 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim DOI: 10.1002/marc.200900025
Segmental Mobility in the Non-crystalline Regions of . . .

from 133.5 8C to 135.5 8C and a fast melting ranging from


136 8C to 141 8C. To probe the structural changes during the
melting processes in the fast and slow melting regions on
the molecular level 1H NMR studies were performed.

Melting in UHMW-PE Probed by 1H NMR


1
H NMR lineshapes and spin-spin (T2) relaxation are
strongly affected by rotational chain motions on time-
scales < 103 s.[12] The more restricted the motion, the
Figure 10. Illustration of the crystal structure of a) melt-crystal- broader is the resonance and the faster the relaxation decay
lised sample having inter crystal molecules and loose loops. b) of the so-called Hahn-echo produced by a 908-t-1808-t pulse
Nascent disentangled sample having tight folds c) Nascent
sequence versus the echo time 2t. Spin-spin relaxation is a
entangled sample having tight folds. The depicted entangle-
ments in the figure are circled. better measure for polymer chain motion than the spectral
line shape, because the latter is also broadened by other
mechanisms, which are eliminated in the Hahn-echo
experiment. 1H NMR T2 relaxometry is therefore a regular
tool to determine entanglement- and crosslink density in
In the melt-crystallised UHMW-PE sample, on the other
rubbers,[29,30,31,32] and polymers melts.[33,34] At T  Tg chain
hand, chain segments in the non-crystalline region are free
mobility at timescales <103 s is controlled by chemical
to adopt a larger number of chain conformations, the stems
crosslinks and physical entanglements with lifetimes
in the crystal are connected by loose folds and are
>103 s. For simplicity of data analysis it is assumed that
dynamically decoupled. Then, melting requires cooperative
polymer motions can be divided into fast and slow
mobility of one chain stem only, corresponding to the
dynamics compared to 103s without a significant inter-
crystal thickness (shown schematically in Figure 10(a)).
mediate fraction. To distinguish between the crystalline
Therefore, the number of repeating units involved in the
and non-crystalline regions of entangled and disentangled
melting process corresponds to the lamella thickness of the
nascent UHMW-PE, the relatively robust and quantitative
crystal, resulting in a melting point of 135 8C, as predicted
Hahn-echo method as a tool is applied.
from the Gibbs-Thomson equation. 1
H NMR Hahn echo decays were monitored at constant
By modifying the Gibbs-Thomson equation, Höhne[17,28]
temperature (Figure 11 and Figure 12). Between the
correlated the number of CH2 units cooperatively involved
experiments at different temperatures, the sample is
in the melting dynamics with the melting temperature. In
heated at a rate of 0.1 K min1. Below the onset of melting,
this approach the number of CH2 groups of the
the observed Hahn-echo decays of the nascent entangled
respective molecule, incorporated into the crystallite via
UHMW-PE sample are well described in terms of two
tight folds is the essential variable, rather than the crystal
exponential components (Figure 11). The fastest compo-
thickness. The total (molar) enthalpy DmHmol and entropy
nent has a fairly temperature-independent short T2 value
DmSmol of fusion (melting) of chain molecules each
consisting of n repeat units in the crystals is composed of
the respective values of the subunits. Accordingly, in the
modified Gibbs-Thomson equation crystal thickness is
replaced by the number of repeating units n and the total
excess (Gibbs) free energy replaces the surface free energy,
yielding

 
1 1 Dm Ge 1
n
 1
1  1 n
(3)
Tm Tm Dm Hr:u:

In the case of cluster melting of 7–8 stems, with n  1 800


CH2 units, a melting temperature of 139.7 8C is predicted in
agreement with the experimentally determined melting
temperature (138.2 8C) obtained by extrapolation to the
zero heating rate (Figure 2). Figure 6, however, clearly shows Figure 11. T2 relaxation time versus temperature of nascent
two distinct melting regions, i.e., a slow melting ranging entangled UHMW-PE (reproduced from ref. [33]).

Macromol. Rapid Commun. 2009, 30, 826–839


ß 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.mrc-journal.de 835
S. Rastogi, Y. Yao, G. W. H. Höhne, R. Graf, H. W. Spiess, P. J. Lemstra

Figure 12. T2 relaxation time versus temperature of the nascent


Figure 13. Changes in the relative fractions of the crystalline,
disentangled UHMW-PE (reproduced from ref. [33]).
mobile non-crystalline and normal non-crystalline components
of the disentangled nascent UHMW-PE sample with temperature
as probed from a Hahn echo pulse program (reproduced from
ref. [33]).
10 ms and disappears above 140 8C and is assigned to the
rigid, crystalline regions. Above 120 8C, the long T2
relaxation time of the non-crystalline fraction increases
strongly with temperature. The steepest increase occurs that the polymer chain stems that are molten by the
around the melting point Tm  140 8C. Apparently, below successive detachment are able to entangle with the
Tm, chain motion in the entangled non-crystalline region is surrounding non-crystalline region. Upon increasing
confined by the crystalline domains, see also ref. [35]. The the temperature further with the melting of the remainder
resulting melt state is described by a single T2 relaxation of the disentangled crystalline fraction (approximately
time of 1 ms. 40%), a highly mobile non-crystalline component of the
Using a similar heating protocol, a third T2 component same fraction evolves. The amount of the highly mobile
shows up in the Hahn-echo decays of the nascent non-crystalline fraction as well as the normal non-crystal-
disentangled sample (Figure 12) about 2–3 times longer line fraction remains almost constant upon further
than that of the non-crystalline region. Because of its long T2 increasing the temperature. This suggests a barrier for
value, we attribute the third T2 component to a more mobile the mixing of the two non-crystalline-regions. As the
type of non-crystalline region to be distinguished from the polyethylene chains are chemically uniform, the presence
normal non-crystalline region observed for the entangled of the two non-crystalline fractions with two different
and disentangled nascent UHMW-PE. The third T2 mobile mobilities above 140 8C, originate from the ‘heterogeneous’
amorphous component arises on slow melting of the distribution of entanglements, which arises on slow
disentangled polyethylene crystals. melting of the disentangled crystals.[3]
Quantitative analysis of the nascent disentangled On fast heating (10 K  min1) the disentangled nascent
sample is shown in Figure 13. The crystalline fraction sample to its melt state, the T2 relaxation decay determined
(shortest T2 component) of approximately 72% as deter- at 150 8C is faster than that of the slowly heated sample
mined from the Hahn-echo decay at 120 8C is consistent (0.1 8C  min1) at the same temperature (see Figure 14(a)).
with the crystallinity of the sample estimated from DSC. As In the fast heated sample, the mobile non-crystalline
the crystallinity decreases with increasing temperature component is absent and the T2-relaxation can be described
from 130 8C to 137 8C, the normal non-crystalline fraction with a single relaxation time. The T2-relaxation at 150 8C
increases (Figure 13). Upon further increasing the tempera- determined for the entangled nascent sample, shown in
ture above 137 8C, the highly mobile non-crystalline region Figure 14(b), is independent of heating rate. Using the
increases simultaneously with the disappearance of the different T2 relaxation time as a filter, thereby suppressing
crystalline region. the less-mobile fractions of the melt, we observe significant
To recall, melting at low temperatures (observed with a differences in the peak width of the NMR spectrum of the
heating rate of 0.1 K  min1 from 130 8C to 137 8C) is two melt states arising with different heating rates from
associated with the successive detachment and diffusion of the initially disentangled nascent sample. The narrower
a few chain stems from the crystal surface (Figure 6). The peak of the slowly heated melt indicates a higher local
increase of the normal non-crystalline fraction suggests mobility in part of the sample. In fact, an increase in T2 filter

Macromol. Rapid Commun. 2009, 30, 826–839


836 ß 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim DOI: 10.1002/marc.200900025
Segmental Mobility in the Non-crystalline Regions of . . .

Melt Mechanism

Combining the findings from DSC and


1
H NMR a scheme for the melt mechan-
ism involved in melting of the disen-
tangled nascent UHMW-PE was devel-
oped. Figure 15(a) depicts slow melting of
the disentangled nascent crystals. The
adjacently re-entrant chains melt by
consecutive detachment of a chain seg-
ment (stem) equal to the lamella crystal
thickness from the crystal lattice (AA’and
BB’in Figure 15(a)). The initially disen-
tangled molten chains will form entan-
glements by reptation (circles) leading to
a shorter T2 relaxation time. Upon melt-
ing the whole crystal at higher tempera-
tures, the section AB in Figure 15(a), will
transform from a disentangled crystal to
a disentangled non-crystalline region (as
was shown in the NMR experiments). As
a result of the absence of physical
entanglements, the disentangled com-
ponent of the non-crystalline fraction
can be described by a high T2 value
referring to the overall higher local
mobility. Thus, depending on the heating
rate, chain dynamics in the resultant
melt state of the disentangled nascent
crystals can be altered. Detailed study on
the influence of heating rate on chain
dynamics is reported elsewhere.[3,36]
Figure 14. NMR line shapes and T2 relaxation curves of fast (10 K  min1, filled squares)
Contrary to the slow melting, a com-
and slow (0.1 8C  min1, unfilled square) heated UHMW-PE melts of the initially
entangled and disentangled nascent samples. Experiments were performed at 150 8C. plete collapse of the crystal lattice occurs
The dynamic heterogeneity of the slow heated disentangled nascent sample can be upon fast melting. This causes an instant
monitored as well via variation of the line width in the T2-filtered spectra. (10 ms is the free movement of the chain ends in the
applied T2 relaxation filter time). Reproduced from ref. [33]. melt, resulting in a homogenous distri-
bution of entanglements (Figure 15(b)).
time leads to suppression of the rigid fractions of the sample Finally, in the nascent entangled samples, physical
and decreasing peak width in the slow heated disentangled entanglements present in the non-crystalline region of the
melts, strengthening the idea of heterogeneity in the local semicrystalline polymer restrict the consecutive detach-
mobility.[3] ment of chains. Therefore the low activation energy
These dynamic NMR results suggest that upon fast component (slope (a) in Figure 6) is absent. Thus on
melting, associated to melting in clusters of 7–8 polymer melting, independent of the heating rate, entanglements
chain stems, the chains are homogenously distributed in present in the non-crystalline region become homoge-
the melt, resulting in a homogenous distribution of neously distributed along the main chain (Figure 15(c)).
entanglements. If the disentangled nascent UHMW-PE is A theory on the evolution of the heterogeneous melt state
given more time to melt (slow heating), the crystals are arising on the controlled melting of the disentangled
molten first from the sides (chain ends) before complete crystals is proposed by McLeish.[36] The influence of the
breakdown of the crystal lattice occurs. The resulting mobile disentangled domains on crystallisation kinetics is
melt state contains a heterogeneous distribution of addressed in ref. [37] Molar mass dependence on build-up of
chain entanglements. The resulting melt state exhibits the modulus on melting of the disentangled samples as a
interesting rheological phenomena as summarised in function of time is addressed in ref. [38] Details of the work
ref.[3,37] can be found in a PhD thesis.[39]

Macromol. Rapid Commun. 2009, 30, 826–839


ß 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.mrc-journal.de 837
S. Rastogi, Y. Yao, G. W. H. Höhne, R. Graf, H. W. Spiess, P. J. Lemstra

scale. Melting in the low temperature


region (130–135 8C) occurs by cooperative
detachment of chain stems from the
surface involving a time-dependent pro-
cess, whereas melting in the high tem-
perature region (136–141 8C) leads to a
breakdown of larger parts of the lattice.
Therefore the normal extrapolation of the
observed Tm to zero heating rate of
Figure 2 needs modification at low
heating rates, as shown in Figure 16.
Recently, similar to the disentangled
nascent polyethylenes Toda et al. have
also attributed the melting kinetics in the
single crystals of linear polyethylenes to
an entropic barrier.[18]
The high melting temperature of nas-
Figure 15. Depicting melting process of the disentangled nascent crystals during slow
heating. (a) On annealing below 137 8C, with the consecutive detachment of chain stems
cent and solution-crystallised UHMW-PE
from the crystal substrate and their diffusion in the melt, normal entangled non- compared to that of melt-crystallised
crystalline region is formed. On heating above 137 8C, the remainder of the crystal melts UHMW-PE is ascribed to the differences
invoking the mobile non-crystalline region. The two non-crystalline regions, mobile and in the chain topology. Depending on the
normal, do not mix even above the melting temperature. This leads to the origin of a nature of chain folds in the crystal, loose
melt having differences in the local mobility of the two non-crystalline components.
Entanglements are encircled. (b) Depicting the melting process of the disentangled
or tight, the number of CH2 units that
nascent crystals during fast heating. (c) Depicting the melting process of the entangled require cooperative motion for melting
nascent crystals. On melting, entanglements initially present in the non-crystalline differs. Melt-crystallised samples with
region get homogeneously distributed. Reproduced from ref. [3]. loose folds melt at the temperature
predicted by the Gibbs-Thomson equa-
tion, i.e., the number of CH2 units that
require cooperative motion to adopt a random coil
corresponds to the crystal thickness. In nascent and
solution-crystallised samples, however, having tight folds,
the number of CH2 units required for cooperative melt
dynamics is much larger than calculated from the crystal
thickness. Therefore, the melting point of the samples shifts
to higher temperatures than in the melt-crystallised
samples.
1
H NMR shows the influence of the different melting
processes on the local mobility of the non-crystalline
regions and in the melt. The chain ends of the disentangled
nascent sample, which detach from the crystal sides prior to
melting (T < 135 8C) form entangled non-crystalline
Figure 16. The measured melting peak temperatures for the domains. Melting of the remainder of the crystal occurs
entangled and the disentangled nascent UHMW-PE at different
at a higher temperature, leading to a melt with mobile
heating rates. The non-linear abscissa allows extrapolation to
zero heating rate0. However, the extrapolation fails at the lower disentangled and normal entangled non-crystalline
heating rates. domains. With the help of 1H NMR, a remarkable distinction
in the distribution of the topological constraints in the melt
can be probed. The disentangled melt state is consistent
Conclusion with the lower plateau modulus determined by rheological
studies.[3]
From the series of experiments reported above it is evident
that melting of nascent-, solution-crystallised and melt-
crystallised UHMW-PE samples shows profound differ- Acknowledgements: Financial assistance from the Dutch Polymer
ences. The number of CH2-groups that are involved in a melt Institute and DSM Research (The Netherlands) is gratefully
process depends on the temperature as well as on the time acknowledged.

Macromol. Rapid Commun. 2009, 30, 826–839


838 ß 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim DOI: 10.1002/marc.200900025
Segmental Mobility in the Non-crystalline Regions of . . .

Received: January 13, 2009; Revised: March 23, 2009; Accepted: [18] A. Toda, M. Hikosaka, K. Yamada, Polymer 2002, 43, 1667.
March 24, 2009; DOI: 10.1002/marc.200900025 [19] From T. Y. Cho, B. Heck, G. Strobl, Colloid Polym Sci. 2004, 282,
Keywords: differential scanning calorimetry (DSC); lamellar; 825, the calculated crystal thickness for the melting tempera-
melting point; NMR; polyethylene (PE) ture of 138.4 8C is 47 nm; this is also larger than the measured
thickness.
[20] Y. F. Yao, R. Graf, H. W. Spiess, D. R. Lippits, S. Rastogi, Phys.
[1] P. Smith, H. D. Chanzy, B. P. Rotzinger, Polym. Commun. 1985, Rev. E 2007, 76(6), 060801.
26, 258. [21] Y. F. Yao, R. Graf, H. W. Spiess, S. Rastogi, Macromolecules
[2] B. P. Rotzinger, H. D. Chanzy, P. Smith, Polymer 1989, 30, 1814. 2008, 41, 2514.
[3] S. Rastogi, D. R. Lippits, G. W. M. Peters, R. Graf, Y. Yao, H. W. [22] D. L. Vanderhart, J. Magn. Resonan. 1981, 44, 117.
Spiess, Nature Mat. 2005, 4, 635. [23] Y. F. Yao, R. Graf, H. W. Spiess, S. Rastogi, Macromol. Rapid
[4] P. J. Lemstra, C. W. M. Bastiaansen, S. Rastogi, ‘‘Structure Commun. 2009, 30, in press.
Formation in Polymeric Fibers’’, Chapter 5, D. R. Salem, Ed., [24] P. W. Atkins, ‘‘Physical Chemistry’’, 5th Edition, Chapter 24.6, 27.4,
Hanser, Munich 2000. 28.12, Oxford University Press, 1995.
[5] S. Rastogi, A. B. Spoelstra, J. G. P. Goossens, P. J. Lemstra, [25] Considering 1/3rd lesser neighbor interactions on the surface
Macromolecules 1997, 30, 7880. than in bulk, the detachment energy and its diffusion into
[6] Y. M. T. Tervoort-Engelen, P. J. Lemstra, Polym. Commun. 1991, the melt is likely to be 2.7 kJ  mol1 since the melting
32, 343. enthalpy of the bulk is 4.11 kJ  mol CH1 2 , a value obtained
[7] S. Rastogi, L. Kurelec, D. R. Lippits, J. Cuijpers, M. Wimmer, P. J. from the ATHAS data bank (http://athas.prz.rzeszow.pl/).
Lemstra, Biomacromolecules 2005, 6, 942. [26] B. Wunderlich, Prog. Polym. Sci. 2002, 391, 51.
[8] L. Corbeij-Kurelec, ‘‘Chain mobility in polymer systems’’, Chap- [27] G. Strobl, ‘‘The Physics of Polymers’’, 2nd edition, Springer-
ter 3, Ph.D. thesis, Eindhoven University of Technology 2001; Verlag, Berlin 1997.
http://alexandria.tue.nl/extra2/200113706.pdf [28] G. W. H. Höhne, Thermochim. Acta 2003, 403, 25.
[9] B. Wunderlich, G. Czornyj, Macromolecules 1977, 10, 906. [29] C. G. Fry, A. C. Lind, Macromolecules 1988, 21, 1292.
[10] The authors are aware that depending on the experimental [30] M. E. L. Wouters, V. M. Litvinov, F. L. Binsbergen, J. G. P.
methods used, different numerical Gibbs-Thomson Goossens, M. van Duin, H. G. Dikland, Macromolecules
equations exist (see: T. Y. Cho, B. Heck, G. Strobl, Colloid 2003, 36, 1147.
Polym. Sci. 2004, 282, 825). A difference arises because of [31] D. L. Tillier, J. Meuldijk, P. C. M. M. Magusin, A. M. van Herk,
different surface free energy values resulting in a somewhat C. E. Koning, J. Polym. Sci. A, 2005, 43, 3600.
different melting temperature of 136 8C for a crystal thick- [32] R. A. Orza, P. C. M. M. Magusin, V. M. Litvinov, M. van Duin,
ness of 25 nm. However, such discrepancies in the calculated M. A. Michels, Macromol. Symp. 2005, 230, 144.
melting temperatures have no implications on our exper- [33] T. Cosgrove, M. J. Turner, P. C. Griffiths, J. Hollingshurst, M. J.
imental findings. Shenton, J. A. Semlyen, Polymer 1996, 37, 1535.
[11] D. R. Lippits, S. Rastogi, G. W. M. Höhne, Phys Rev Lett. 2006, 96, [34] A. Guillermo, J. P. Cohen Addad, D. Bytchenkoff, J. Chem. Phys.
218303. 2000, 113, 5098.
[12] K. Schmidt-Rohr, H. W. Spiess, ‘‘Multidimensional solid-state [35] H. W. Spiess, Colloid Polym. Sci. 1983, 261, 193.
NMR and polymer’’, Academic Press Ltd, San Diego, USA 1994. [36] T. C. B. McLeish, Soft Matter 2007, 3, 83.
[13] K. Saalwächter, I. Schnell, Solid State Nucl. Magn. Reson. 2002, [37] D. R. Lippits, S. Rastogi, S. Talebi, C. Bailly, Macromolecules
22, 154. 2006, 39, 8882.
[14] K. Schmidt-Rohr, H. W. Spiess, Macromolecules 1991, 24, 5288. [38] D. R. Lippits, S. Rastogi, G. W. M. Höhne, B. Mezari, P. C. M. M.
[15] G. W. H. Höhne, L. Kurelec, S. Rastogi, P. J. Lemstra, Thermo- Magusin, Macromolecules 2007, 40, 1004.
chim. Acta 2003, 396, 97. [39] D. R. Lippits, ‘‘Heterogeneity in polymer melts by controlled
[16] G. W. H. Höhne, Thermochim. Acta 1997, 304/305, 209. melting of polymer crystals’’, PhD Thesis, Eindhoven Univer-
[17] G. W. H. Höhne, Polymer 2002, 43, 4689. sity of Technology, March 6th 2007.

Macromol. Rapid Commun. 2009, 30, 826–839


ß 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.mrc-journal.de 839

You might also like