Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

Polymer 54 (2013) 1709e1728

Contents lists available at SciVerse ScienceDirect

Polymer
journal homepage: www.elsevier.com/locate/polymer

Feature article

Novel polymer ferroelectric behavior via crystal isomorphism and the


nanoconnement effect
Lianyun Yang a, Xinyu Li b, Elshad Allahyarov c, d, e, Philip L. Taylor c, Q.M. Zhang b, **, Lei Zhu a, *
a

Department of Macromolecular Science and Engineering, Case Western Reserve University, Cleveland, OH 44106-7202, USA
Department of Electrical Engineering, Pennsylvania State University, University Park, PA 16802, USA
c
Department of Physics, Case Western Reserve University, Cleveland, OH 44106-7079, USA
d
Institut fr Theoretische Physik, Heinrich-Heine-Universitt Dsseldorf, D-40225 Dsseldorf, Germany
e
Theoretical Department, Joint Institute for High Temperatures, Russian Academy of Sciences, Izhorskaya 13/19, 117419 Moscow, Russia
b

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 29 November 2012
Received in revised form
13 January 2013
Accepted 19 January 2013
Available online 4 February 2013

In contrast to the comprehensive understanding of novel ferroelectric [i.e., relaxor ferroelectric (RFE) and
antiferroelectric] behavior in ceramics, RFE and double-hysteresis-loop (DHL) behavior in crystalline
ferroelectric polymers have only been studied in the past fteen years. A number of applications such as
electrostriction, electric energy storage, and electrocaloric cooling have been realized by utilizing these
novel ferroelectric properties. Nonetheless, fundamental understanding behind these novel ferroelectric
behaviors is still missing for polymers. In this feature article, we intend to unravel the basic physics via
systematic studies of poly(vinylidene uoride-co-triuoroethylene) [P(VDF-TrFE)]-based terpolymers,
electron-beam (e-beam) irradiated P(VDF-TrFE) copolymers, and PVDF graft copolymers. It is found that
both the crystal internal structure and the crystaleamorphous interaction are important for achieving
the RFE and DHL behaviors. For the crystal internal structure effect, dipole switching with reduced
friction and nanodomain formation by pinning the polymer chains are essential, and they can be achieved through crystal repeating-unit isomorphism (i.e., defect modication). Physical pinning [e.g., in
P(VDF-TrFE)-based terpolymers] induces a reversible, electric eld-induced RFE4FE phase transition
and thus the DHL behavior, whereas chemical pinning [e.g., in e-beam irradiated P(VDF-TrFE)] results in
the RFE behavior. Finally, the crystaleamorphous interaction (or the nanoconnement effect) results in
a competition between the polarization and depolarization local elds. When the depolarization eld
becomes stronger than the polarization eld, a DHL behavior is observed. Obviously, the physics for
ferroelectric polymers is different from that for ceramics/liquid crystals and can be largely attributed to
the long-chain nature of semicrystalline polymers. This understanding will help us to design new ferroelectric polymers with improved properties and/or better applications.
2013 Elsevier Ltd. Open access under CC BY-NC-ND license.

Keywords:
Crystal isomorphism
Nanoconnement effect
Relaxor ferroelectric behavior

1. Novel ferroelectric polymers with different nonlinear


dielectric properties
Different from ceramic ferroelectrics, ferroelectric (FE) polymers
represent a unique class of functional insulating materials due to
their long-chain rather than ionic-crystal nature. The advantages of
easy processing and low cost have made them attractive for a variety
of practical applications. By denition, ferroelectricity refers to both
structure and property aspects. Structure-wise, ferroelectric materials must possess polar structures with spontaneous polarization.
Property-wise, ferroelectric materials exhibit broad hysteresis loops
* Corresponding author. Tel.: 1 216 368 5861.
** Corresponding author. Tel.: 1 814 863 8994.
E-mail addresses: qxz1@psu.edu (Q.M. Zhang), lxz121@case.edu (L. Zhu).
0032-3861 2013 Elsevier Ltd. Open access under CC BY-NC-ND license.
http://dx.doi.org/10.1016/j.polymer.2013.01.035

due to the existence of large spontaneous polarization (Fig. 1A and


E). Here, D, the electric displacement, is the total charge density
induced by applying an electric eld (E) on a dielectric material,
and is dened as D E. In this equation, is the permittivity and is
dened as r0, where r is the relative permittivity (or dielectric
constant) and 0 is the permittivity of the vacuum. For linear dielectrics, the r is a constant. For nonlinear dielectrics, the r is
a function of the electric eld. Generally speaking, the electrical
relationship between D and E is similar to the mechanical relationship between strain and stress, respectively. It is this normal ferroelectricity that makes ferroelectric polymers attractive for many
applications such as nonvolatile memory [1e6] and transducers in
sensors and actuators [7e12]. In the other extreme case, when no
ferroelectric domains exist [e.g., in nonpolar polyethylene/polypropylene or when dipoles cancel each other out, as in even-

1710

L. Yang et al. / Polymer 54 (2013) 1709e1728

Fig. 1. Schematic hysteresis loops for (A) normal ferroelectric, (B) paraelectric or dipolar glass or relaxor ferroelectric, (C) antiferroelectric, and (D) dielectric materials. From (D) to
(A) the dielectric nonlinearity gradually increases. The bottom panel illustrates the ferroelectric domain structures of (E) normal ferroelectric, (F) relaxor ferroelectric, (G) dipolar
glass, and (H) dielectric polymers. From (H) to (E) the domain size gradually increases.

numbered nylons], the polymers display linear electric polarization


(i.e., D is a linear function as E; see Fig. 1D and H) [13]. The origin of
the dielectric behavior lies in a simple fact that only electronic and
atomic polarizations, rather than dipolar orientational, ionic, and
interfacial polarizations, work in the dielectric materials. In between
these two extreme cases, paraelectricity refers to the ability of polar
materials with permanent dipoles to become polarized under an
electric eld, but upon removal of the applied eld the polarization
quickly returns to zero (the loop is similar to Fig. 1B) [13,14]. From
a structure point of view, paraelectric (PE) polymers should contain
permanent dipoles but without any ferroelectric domains at zero
electric eld (similar to the case of dielectrics in Fig. 1H).
Recently, novel ferroelectric behaviors have emerged for ferroelectric polymers and have become increasingly attractive for new
applications such as electrostriction [9,15e26], electric energy storage [13,27e40], and electrocaloric cooling [41e53]. These include
narrow single (Fig. 1B) and double (Fig. 1C) hysteresis loop behaviors.
For ceramics, these ferroelectric behaviors are not new and have
been relatively well-understood. For example, it is known that
decreasing the ferroelectric domain size to accommodate a few dipoles [i.e., nanodomains in relaxor ferroelectrics (RFE)] or just one
dipole (i.e., dipolar glasses) will signicantly mitigate the cooperative coupling among ferroelectric domains, resulting in a reduction
of the spontaneous polarization and thus narrow single hysteresis
loops [54,55]. For antiferroelectric (AFE) ceramics, double-hysteresis-loops (DHLs) are observed due to reversible AFE4FE transitions
as the applied electric eld reverses direction [56]. Nonetheless,
these novel RFE and AFE ferroelectric behaviors are still rare (e.g., the
RFE phase) or nonexistent (e.g., the AFE phase) in crystalline polymers. This can be attributed to the long-chain nature of ferroelectric
crystalline polymers. For example, nanodomain formation in closepacked long-chain polymer crystals is more difcult to achieve
than that in compositionally disordered perovskite (ABO3) crystals
composed of different ions on crystallographically equivalent sites,
because ferroelectricity is achieved by 180 -ipping of permanent
dipoles in the main-chain of ferroelectric polymers whereas ferroelectricity can be easily achieved by displacement of the center ion
towards eight corners of the tetragonal crystal structure of ceramic
ferroelectrics. As a result, the RFE behavior has only been reported

for electron-beam (e-beam) or g-ray irradiated poly(vinylidene


uoride-co-triuoroethylene) [P(VDF-TrFE)] [16,17,22,25,26,57] and
P(VDF-TrFE)-based terpolymers [21,22,25,58e64]. The long-chain
nature of polymer crystals also makes the occurrence of the
AFE/FE transition difcult, if not impossible. For example, in evennumbered nylons the hydrogen-bonded amide dipoles are arranged
in an AFE-like structure with an alternating up and down dipolar
conguration in a long chain due to the zig-zag conformation of the
alkane spacers [65,66]. It is impossible for every other hydrogenbonded amide dipole to ip 180 and form a polar arrangement in
a close-packed crystal. Therefore, it is important to understand the
fundamental physics behind these novel ferroelectric behaviors for
semicrystalline polymers.
Fundamentally, most ferroelectric polymers are semicrystalline
(note that we will not focus on ferroelectric liquid crystalline polymers in this feature article). For example, typical crystallinity for
PVDF and its copolymers with chlorotriuoroethylene (CTFE) and
hexauoropropylene (HFP) are below 50 wt.% primarily because of
structural defects such as head-to-head/tail-to-tail (HHTT) sequence
(usually 3e6 mol% by conventional radical polymerization) and/or
comonomers in the main chain [67]. For P(VDF-TrFE) and poly(VDFco-tetrauoroethylene) [P(VDF-TFE)] random copolymers, repeating-unit isomorphism [65,68] takes place, because TrFE and TFE are
similar in both chemical structure and size as VDF. Consequently, the
crystallinity signicantly increases to >80 wt.%. Sometimes, when
annealed in the PE phase, the crystallinity of P(VDF-TrFE) can reach
nearly 100 wt.% [69e71]. For these semicrystalline polymers, it is
understandable that their ferroelectric properties are intimately
dependent upon both the internal crystal structure and the crystale
amorphous interaction (or the nanoconnement effect [72,73]).
Generally speaking, the crystalline structure is a major factor that
affects the ferroelectric behavior of polymers, whereas the nanoconnement effect is a less important factor.
2. Effect of crystal isomorphism on novel ferroelectric
behaviors of P(VDF-TrFE)-based copolymers
Before discussing ferroelectric behavior in P(VDF-TrFE)-based
random copolymers, it is necessary for us to understand the

L. Yang et al. / Polymer 54 (2013) 1709e1728

normal ferroelectric behavior in terms of close-packing (or the area


packing density in the ab-plane; note that the molecular chain direction is along the c-axis) in PVDF crystals. For a b PVDF crystal, the
average area packing density in the ab-plane is 82.6%, signicantly
higher than that (70%) for a nonpolar polyethylene crystal. As a zigzag PVDF stem (contains perpendicular dipoles) inside a ferroelectric domain ips (at every rotation step of ca. 60 [74,75]) in
response to an alternating eld, it must experience signicant
intermolecular friction from its neighbors. We consider that it is the
strong intermolecular friction and large spontaneous polarization
that prevent the easy rotation of PVDF chains inside the closepacked crystal. To achieve easy rotation of zig-zag PVDF chains, it
is highly desirable to reduce intermolecular friction (by expanding
the interchain distance) and to decrease spontaneous polarization
(by reducing the ferroelectric domain size).
2.1. PE behavior in P(VDF-TrFE) random copolymers
As mentioned above, P(VDF-TrFE) random copolymers demonstrate type-2 repeating-unit isomorphism [7], because the two
homopolymers have different crystal structures (PVDF has four
orthorhombic crystal structures (see Fig. 4 in Ref. [13]) and PTrFE
has a hexagonal crystal structure [76]). Consequently, both the
interchain distance and overall crystallinity increase as compared
with the homopolymers. For example, the (110/200) d-spacing
A for the low temperature FE phase in a P(VDF(d110/200) is 4.42 
TrFE) 50/50 (mol/mol) random copolymer, and it is larger than
A for the b PVDF crystal. However, the average
the d110/200 of 4.26 
area packing density in the ab-plane for the low temperature FE
phase is 78.0%, not much lower than that (82.6%) for PVDF. Therefore, normal ferroelectric behavior with a broad hysteresis loop is
still observed for the low temperature FE phase, although switching
of zig-zag P(VDF-TrFE) chains becomes easier. The d110/200 for the
high-temperature PE phase of P(VDF-TrFE) 50/50 increases to
4.86 
A, and the average area packing density in the ab-plane decreases to ca. 70%, signicantly lower than those for both b PVDF
and low temperature FE P(VDF-TrFE) crystals. As a result, the dipole
and chain ipping become much easier for the high-temperature
PE phase of P(VDF-TrFE). This is exactly observed in a DeE loop
study in a recent report (see Fig. 2) [77]. At 100  C [above the Curie
temperature (TC) at 64  C], narrow single loops with minimum
hysteresis are observed for the high-temperature PE phase at
a poling frequency of 1000 Hz (Fig. 2D). This is drastically different
from the open loops observed for the low temperature FE phase at
a poling frequency of 1000 Hz at 25  C (Fig. 2A).
Moreover, the ferroelectric behavior of the high-temperature PE
phase is dependent upon poling frequency (as is the low temperature FE phase). The hysteresis loops gradually open up as the
poling frequency decreases from 1000 Hz to 1 Hz (Fig. 2DeF). This is
somewhat inconsistent with the denition of the PE phase; however, it can be explained by the schematic illustration in Fig. 2G.
Below the TC, large ferroelectric domains having a large spontaneous polarization exist in the sample (Fig. 2G a). Above the TC,
these ferroelectric domains completely disappear with randomly
mixed TG (T is trans and G is gauche) and TTTG conformations in
P(VDF-TrFE) chains (Fig. 2G b). However, applying a high enough
electric eld will induce both nucleation and growth of ferroelectric
domains (Fig. 2G cee). This is conrmed by the eld-dependent
Fourier transform infrared (FTIR) results in Fig. 2H. Below the TC,
the long T sequence in the FE phase is evidenced by the absorption
peak at 507 cm1. Above the TC, the mixed TG and TTTG conformation in the PE phase is evidenced by a broadband at 514 cm1.
Upon application of an electric eld of 50 MV/m, a shoulder starts
to appear at 507 cm1, suggesting the nucleation and growth of
ferroelectric domains with a long T sequence in the chain. On

1711

further increasing the electric eld to 125 MV/m, the 507 and
514 cm1 bands merge into one centered at 510 cm1. As shown in
Fig. 2G, the high-frequency poling induces nucleation of nanosized
ferroelectric domains (nanodomains) within the weakly interactive
paraelectric matrix (Fig. 2G e). Consequently, narrow single hysteresis loops with minimal spontaneous (or remanent) polarization
are achieved. Upon decreasing the poling frequency to 1 Hz,
nanodomains grow into relatively large ones (Fig. 2G d and c), and
thus open ferroelectric loops with relatively large spontaneous (or
remanent) polarization are observed.
Frequency-dependent dipole switching is also reected in the
broadband dielectric spectroscopy (BDS) results for a uniaxially
stretched P(VDF-TrFE) 50/50 lm during the rst and second heating
cycles under different frequencies (Fig. 3). During the rst heating
cycle (Fig. 3A and B), the glass transition temperature (Tg) is observed
around 20  C at 1 Hz and it increases with increasing poling frequency. Meanwhile, a broad Curie transition is observed for 1 Hz,
starting around 35  C and showing a relatively sharp peak at 80  C.
This result corresponds well with the differential scanning calorimetry (DSC) result in Fig. 3H, where two Curie transitions are seen
at 50 and 80  C, respectively. After heating to 100  C and cooling back
to 50  C, the history of mechanical stretching on the Curie transition is removed. Subsequent BDS results during the second heating
cycle are shown in Fig. 3C and D. In addition to the frequencydependent glass transition peak, a relatively sharp Curie transition
peak is observed at 65  C for 1 Hz and it gradually broadens and
shifts to higher temperatures upon increasing the poling frequency.
Comparing the BDS results during the rst and second heating cycles, we infer that the TC must be domain size dependent. On one
hand, uniaxial stretching enhances the crystal orientation with
enlarged domain size in some P(VDF-TrFE) crystallites, because
a higher TC is observed at 80  C relative to 64  C seen during the
second heating (see Fig. 3H). On the other hand, some crystallites are
torn apart with a decreased domain size, and thus the TC decreases to
50  C correspondingly. A similar domain size dependence for TC has
also been reported in the literature [67,71,78]. Namely, large domains
resulted in a higher TC and small domains resulted in a lower TC. The
activation energies (Ea) for both amorphous (from the glass transition in Fig. 3D) and crystalline (from the Curie transition in Fig. 3C)
dipole relaxations can be obtained from Arrhenius plots of log(peak
frequency) versus 1/T (see Fig. 3E). From the linear regions where the
Arrhenius equation applies, the activation energies for the amorphous and crystalline dipoles are found to be 0.57 and 7.4 eV,
respectively. Obviously, it is much more difcult to switch crystalline
dipoles than amorphous ones for semicrystalline ferroelectric polymers. The nonlinear part for the amorphous dipoles in Fig. 3E can be
explained by the reduced mobility near Tg (i.e., the WLF equation
applies instead), whereas the high-frequency nonlinearity for the
crystalline dipoles may be attributed to the fact that crystalline dipoles cannot follow the electric eld at such high frequencies.
From Fig. 3AeD, upturns in both r 0 and r 00 are observed at
high temperatures (>80  C) and low frequencies. These r 0 and r 00
upturns can be explained by enhanced loss from impurity ions in
the sample (i.e., from suspension polymerization) as temperature
and poling time increase, since the slope in the double-log plot of
the r 00 at low frequencies is nearly 1 [79]. In a recent report, we
demonstrate that a parts-per-million (ppm) or even sub-ppm level
of impurity ions in PVDF can cause signicant increases in both
r 0 and r 00 [80]. Evidence of impurity ion relaxation is demonstrated in r 0 and r 00 as a function of frequency at different temperatures (see Fig. 3F and G). At low temperatures, there is no
upturn in r 0 and only a weak peak in r 00 at low frequencies. On
increasing temperature, the weak peak in r 00 gradually increases its
intensity and shifts to higher frequencies. Finally, above 75  C the
low-frequency loss in r 00 becomes so large that an increase in r 0 is

1712

L. Yang et al. / Polymer 54 (2013) 1709e1728

Fig. 2. Bipolar hysteresis DeE loops for a uniaxially stretched P(VDF-TrFE) 50/50 lm (draw ratio 300%) in the PE and FE phases at 100 and 25  C, respectively: (A and D) 1000 Hz,
(B and E) 10 Hz, and (C and F) 1 Hz. The poling eld (a sinusoidal waveform) starts at 25 MV/m and increases at increments of 25 MV/m. (G) Schematic illustration of the
ferroelectric-to-paraelectric Curie phase transition and electric eld-induced ferroelectric domain nucleation/growth and corresponding ferroelectric behaviors in the paraelectric
phase. Dark blue arrows represent large dipoles in the zig-zag stems in the ferroelectric crystal (light blue rectangle) or domains (light blue ovals). Orange double arrows represent
weak dipoles in the paraelectric crystal (light green rectangles). (H) FTIR spectra of ferroelectric and paraelectric P(VDF-TrFE) 50/50 under different electric elds at 24 and 100  C,
respectively. (Reproduced and adapted with permission from Ref. [77].)

also observed. Meanwhile, a broad dipole relaxation is seen at high


frequencies and the peak gradually shifts to higher frequencies
with increasing temperature. From the plot of log(peak frequency)
versus 1/T obtained from r 00 in Fig. 3G (data not shown), an Ea of ca.
0.3 eV is determined for the dipoles, which is similar to the value of
0.57 eV obtained for amorphous dipoles in Fig. 3E. Therefore, the
high-frequency relaxation should be attributed to amorphous
dipoles.
2.2. DHL behavior in P(VDF-TrFE)-based terpolymers
On the basis of the high-temperature PE behavior of P(VDFTrFE), we have learned that expanded interchain distance and
nanodomains with reduced spontaneous polarization are critical to
reduce intermolecular friction and speed up dipole ipping in ferroelectric crystals. This knowledge is helpful for the design and

development of novel ferroelectric polymers. To further expand the


interchain distance, an even larger comonomer than TrFE needs to
be incorporated into the P(VDF-TrFE) random copolymers. Possible
candidates are vinyl chloride (VC), 1,1-chlorouoroethylene (CFE),
chlorodiuoroethylene (CDFE), chlorotriuoroethylene (CTFE), and
hexauoropropylene (HFP) [58]. Intriguingly, direct copolymerization of VDF with these comonomers most likely will not
result in the formation of repeating-unit isomorphism because of
their large sizes. For example, CTFE repeating units are mostly
excluded from the PVDF crystals in P(VDF-CTFE) and no expansion
of the interchain distance is observed experimentally [27,73].
Therefore, presence of TrFE in PVDF is important for the third
comonomer to be incorporated in the crystal and further expand
the interchain distance.
In this study, we use a P(VDF-TrFE-CFE) 59.2/33.6/7.2 terpolymer as an example to study the ferroelectric behavior after further

L. Yang et al. / Polymer 54 (2013) 1709e1728

1713

Fig. 3. Broadband dielectric spectroscopy results for a uniaxially stretched P(VDF-TrFE) 50/50 lm (draw ratio 300%) during (A and B) the rst and (C and D) the second heating
cycles, respectively. (A and C) and (B and D) show the real r 0 and imaginary r 00 relative permittivity as a function of temperature at various frequencies. (E) Arrhenius plots (log f
versus 1/T) for amorphous and crystalline dipoles. (F and G) r 0 and r 00 as a function of frequency at different temperatures. (H) DSC rst heating, rst cooling, and second heating
curves for the uniaxially stretched P(VDF-TrFE) 50/50 lm.

enlarging the interchain distance. At this CFE composition (7.2 mol


%), the low temperature FE phase has been successfully converted
into an RFE phase [60]. A hot-pressed lm of this terpolymer is
uniaxially stretched in the RFE phase at 10  C to a draw ratio of
400%. Upon mechanical stretching, a certain amount of FE phase is
developed, and this is reected in the rst heating curve in DSC
(Fig. 4) and the two-dimensional (2D) wide-angle X-ray diffraction
(WAXD) pattern in Fig. 5A. Note that the existence of an FE phase in
ultrathin P(VDF-TrFE-CFE) LangmuireBlodgett (LB) lms is also
reported previously [81]. From Fig. 4, two TCs are observed with
a weak peak at 27  C and another peak at 49  C. After meltrecrystallization, only one broad TC is seen at 22  C for the second
heating (Fig. 4). The TC at 49  C can be attributed to the Curie

Heat flow (W/g)

exo

6
9 7 C

1st heating
1st cooling
2nd heating

2
1 5 C

2 7 C

4 9 C

2 2 C

P(VDF-TrFE-CFE) 59.2/33.6/7.2

-2
-5 0

-2 5

25

50

75

1 2 3 C

100

125

150

Temperature (C)
Fig. 4. DSC rst heating, rst cooling and second heating curves for a uniaxially
stretched P(VDF-TrFE-CFE) 59.2/33.6/7.2 terpolymer lm. The stretching is performed
at 10  C and the draw ratio is 400% with a drawing rate of 12 mm/min. The heating and
cooling rates are 10  C/min.

transition for the stretch-induced FE phase and the low temperature TC is attributed to the Curie transition for the RFE phase. Note
that mechanical stretching results in a larger ferroelectric domain
size and thus a slightly higher TC at 27  C for the terpolymer.
Crystal orientations in the uniaxially stretched P(VDF-TrFECFE) terpolymer lm during the rst and second heating cycles
are shown in Fig. 5. From the 2D WAXD pattern in Fig. 5A, both
(110/200) reections from the stretch-induced FE and RFE phases
are seen on the equator and the (001)FE and (002)RFE reections
are seen on the meridian. The corresponding one-dimensional
(1D) WAXD proles on both equator and meridian during the
rst heating cycle are shown in Fig. 5C. At 38  C, the (110/
A, which is similar to the (110/
200)RFE reection is seen at 4.81 
A) of the P(VDF-TrFE) 50/50 [77]. The (110/
200)PE reection (4.86 
A. As the temper200)FE reection appears as a shoulder at 4.57 
ature increases to 47  C, both (110/200)FE and (001)FE reections
disappear, which corresponds to the DSC results in Fig. 4. The
(110/200)RFE reection continuously shifts to lower q values and
A at 90  C. After
nally the (110/200)PE d-spacing increases to 4.98 
the rst heating cycle, the stretch-induced FE phase disappears
and 1D WAXD proles during the second heating cycle are shown
in Fig. 5D. Upon increasing the temperature, the (110/200)
reection continuously shifts to lower q values, and correspondA at
ingly the d-spacing increases from 4.84 
A at 38  C to 4.98 
90  C. Regardless of the (110/200) d-spacing changes, the (002)
A). In
reection remains constant at 27.4 nm1 (d-spacing of 2.29 
contrast with the DSC study, there is no obvious transition for the
(110/200) [and the (002)] reections during the RFE/PE Curie
transition in the second heating cycle (see Fig. S1A in the Supplementary material). We consider that the heating rate of 10  C/
min is so fast that the transition cannot be observable in X-ray
diffraction.
In addition to DSC and WAXD, FTIR is another useful technique
to detect chain conformation changes for semicrystalline polymers.
The FTIR results for the uniaxially stretched P(VDF-TrFE-CFE) terpolymer lm during the rst and second stepwise heating cycles
are shown in Fig. 6. From the rst heating cycle (Fig. 6A), the long T
sequence for the stretch-induced FE phase is evidenced by absorption bands at 507, 848, and 1287 cm1 [82,83]. Note that all

1714

L. Yang et al. / Polymer 54 (2013) 1709e1728

Fig. 5. 2D WAXD patterns of (A) as-stretched and (B) 90  C-annealed uniaxially stretched P(VDF-TrFE-CFE) terpolymer lms (draw ratio 300%). The bottom panel shows 1D WAXD
proles for the equator and meridian reections during (C) the rst and (D) the second heating cycles. The heating rate is 10  C/min and each curve is 5  C apart with the last
temperature being 90  C. The inner two weak and diffuse rings belong to the Kapton tapes used to encapsulate the sample.

Fig. 6. FTIR spectra for the uniaxially stretched P(VDF-TrFE-CFE) terpolymer lm during (A) the rst and (B) the second heating cycles. The spectra are collected during a stepwise
heating process with steps of 5  C. All curves are overlaid for clarity.

L. Yang et al. / Polymer 54 (2013) 1709e1728

1715

Fig. 7. (A and C) Real r 0 and (B and D) imaginary r 00 relative permittivity as a function of temperature under different frequencies for the uniaxially stretched P(VDF-TrFE-CFE)
terpolymer lm during (A and B) the rst and (C and D) the second heating cycles. (E and G) r 0 and (F and H) r 00 as a function of frequency at different temperatures for the
uniaxially stretched P(VDF-TrFE-CFE) terpolymer lm during (E and F) the rst and (G and H) the second stepwise heating cycles.

trans conformation can be obtained for the P(VDF-TrFE-CFE) terpolymers due to the presence of VDF to relieve the steric hindrance
caused by the large Cl atoms in the CFE units. Upon increasing the
temperature to 45  C, the band at 507 cm1 (the second dashed line
from the right) gradually disappears and a new doublet band at
514/527 cm1 appears, indicating the FE/PE transition. The band
at 848 cm1 decreases its intensity and the band at 1287 cm1
disappears. Meanwhile, another doublet band at 605/614 cm1
becomes more pronounced. Note that the absorption bands at 514/
527 and 605/614 cm1 can be assigned to mixed TG and TTTG
conformations [67]. After cooling from 90 to 40  C (Fig. 6B), the
stretch-induced FE phase disappears, because the bands at 507,
848, and 1287 cm1 become weaker and the bands at 514/527 and
605/614 cm1 become stronger. Intriguingly, two bands at 772 and

800 cm1 become much more pronounced, and they are assigned
to the T3G sequence for the g PVDF crystal [84]. Therefore, a short T
sequence (i.e., nanodomains) is resulted for the RFE P(VDF-TrFECFE). This is possibly attributed to the random sequence of CFE
along the chain [i.e., 14 repeating units of VDF/TrFE per CFE unit].
During the second heating cycle (Fig. 6B), although there is no
obvious change for bands at 514/527, 605/616, and 848 cm1 in the
vicinity of the RFE/PE transition at 22  C, the bands at 772 and
800 cm1 clearly decrease their intensities. This indicates the disappearance of nanodomains with a short T3G sequence above the
TC at 22  C.
The dynamic dielectric properties of the uniaxially stretched
P(VDF-TrFE-CFE) terpolymer lm during the rst and second
heating cycles are studied by BDS. As shown in Fig. 7, both the glass

Fig. 8. Bipolar hysteresis loops for the uniaxially stretched P(VDF-TrFE-CFE) terpolymer lm (draw ratio 300%) at (A) 25  C, (B) 0  C, (C) 25  C, (D) 50  C, and (E) 65  C,
respectively. The poling frequency is 10 Hz. Bipolar hysteresis loops under different frequencies at 25  C are also shown: (F) 1000 Hz, (G) 100 Hz, (C) 10 Hz, and (H) 1 Hz. All poling
eld (triangular waveform) starts at 25 MV/m with a stepwise increase of 25 MV/m except for those at 1000 Hz.

1716

L. Yang et al. / Polymer 54 (2013) 1709e1728

and Curie transitions for the amorphous and crystalline dipoles are
seen during heating for r 0 and r 00 as a function of temperature.
For example, during the rst heating cycle, the Tg and TC are
observed at 11 and 39  C (10 Hz), respectively, and they gradually
increase with increasing the frequency (Fig. 7B). During the second
heating cycle (note that the FE phase is eliminated after the rst
heating to 90  C), the Tg and TC are seen at 17 and 12  C (10 Hz),
respectively, and both values increase with increasing the frequency (Fig. 7D). The high TC at 39  C during the rst heating is
attributed to the stretch-induced FE phase, whereas the low TC at
12  C during the second heating is attributed to the RFE phase. From
Fig. 7B and D, we can see that the frequency dependence
(Ea 0.6 eV) of the Tg is stronger than that (Ea 6.7 eV) for the TC,
indicating again that amorphous dipoles are easier to switch than
crystalline dipoles at low electric elds. The high-temperature
(>50  C) and low-frequency (<1 kHz) upturns in the r 00 in Fig. 7B
and D are attributed to the impurity ion relaxation, as revealed by
the frequency scans for the rst and second heating cycles (Fig. 7Ee
H). In Fig. 7E and G, upturns in the r 0 are observed at low

frequencies (<1 Hz) and high temperatures (>50  C). Meanwhile,


in Fig. 7F and H upturns in the r 00 in the double-log plot show
a slope of ca. 1, clearly indicating the loss from impurity ions
[79,80,85]. At high frequencies, a decrease in the r 0 and a broad
peak in the r 00 can be attributed to the relaxation of amorphous
dipoles. At intermediate frequencies, plateaus in the r 0 are
observed in Fig. 7E and G. For the rst heating, the maximum value
in the plateau r 0 is observed for 50  C, whereas during the second
heating it is observed for 25  C. This is because the TC is 49  C for the
stretch-induced FE phase during the rst heating and the TC is 22  C
for the RFE phase during the second heating.
Ferroelectric behavior of the uniaxially stretched P(VDF-TrFECFE) terpolymer lm is studied by DeE hysteresis loop tests. Typical hysteresis loops at different temperatures are shown in
Fig. 8AeE. Below the TC (49  C) of the stretch-induced FE phase,
open loops with a sense of double hysteresis are observed when the
poling eld is above 75 MV/m. At 50  C, the RFE phase has already
transformed into a PE phase and the stretch-induced FE phase
gradually disappears. As a result, slim loops with a weak sense of

Fig. 9. Bipolar hysteresis loops for the 70  C-annealed uniaxially stretched P(VDF-TrFE-CFE) terpolymer lm at (AeD) 25  C, (EeG) 0  C, (IeL) 25  C, and (MeO) 50  C, respectively.
The poling frequency varies from left to right: (A, E, I, M) 1000 Hz, (B, F, J, N) 100 Hz, (C, G, K, O) 10 Hz, and (D, H, L) 1 Hz. Note that at 50  C, the 70  C-annealed lm is easy to
breakdown and thus no results for 1 Hz bipolar poling are obtained. All poling eld (triangular waveform) starts at 25 MV/m with increments of 25 MV/m except for those at
1000 Hz.

L. Yang et al. / Polymer 54 (2013) 1709e1728

double hysteresis are observed above 75 MV/m, similar to the


hysteresis loops for P(VDF-TrFE) at 100  C (see Fig. 1E). When the
temperature increases to 65  C, the center portions of the hysteresis
loops gradually open up with an ellipsoidal shape. This can be
attributed to the enhanced impurity ion loss and electronic conduction at elevated temperatures, consistent with the BDS results
in Fig. 7EeF [80,85]. Frequency-dependent hysteresis loops for the
uniaxially stretched P(VDF-TrFE-CFE) at 25  C are shown in Fig. 8C
and FeH. Upon increasing the poling frequency to 1000 Hz, the
maximum D (Dmax) decreases and loops display more double hysteresis. However, broad loops are still observed because the sample
contains certain FE phase.
After annealing the uniaxially stretched P(VDF-TrFE-CFE) terpolymer lm at 70  C for 1 h, the stretch-induced FE phase disappears and the sample contains 100% RFE phase. Temperaturedependent hysteresis loops at various poling frequencies are shown
in Fig. 9. At 25  C, double loops with signicant hysteresis and
remanent polarization are observed upon decreasing the poling
frequency to below 100 Hz. At 1000 Hz, the crystalline dipoles
cannot follow the electric eld and less hysteresis is seen. The large
hysteresis can be attributed to the larger domain size in the RFE
phase when the temperature is sufciently low. When increasing
the temperature to 0 and 25  C, the hysteresis in the DeE loops
continuously decreases due to a continuous decrease in the domain
size and an increase in the interchain distance. At 25  C, a signicant amount of RFE phase should transform into the PE phase.
Therefore, the remanent polarization reduces to nearly zero and
typical DHLs are observed. The double hysteresis should be caused
by the electric eld-induced (RFE/PE)/FE phase transition upon
poling and the FE/(RFE/PE) phase transition upon removing the
eld, because the FE phase is observed after mechanical stretching.
Currently, we do not have unambiguous evidence of the FE phase
during the in-situ electric poling process. In the future, we will use
time-resolved polarized FTIR [86] to investigate dipole ipping and
phase transitions in uniaxially stretched P(VDF-TrFE-CFE) terpolymer lms. At 50  C (above the TC at 22  C), the sample becomes
100% PE phase. As a result, narrow hysteresis loops are observed at

1717

1000 Hz, similar to the DeE loops at 1000 Hz for the P(VDF-TrFE)
lm in the PE phase (see Fig. 2D). Upon decreasing the frequency,
a sense of double hysteresis is observed at high elds. This again
can be attributed to some FE domain formation within the PE
matrix at high enough electric poling elds.
After annealing the uniaxially stretched P(VDF-TrFE-CFE) lm in
the PE phase at 70  C, the overall crystallinity increases and the lm
appears to be much more brittle. As a result, the lm quality decreases, resulting in a signicant reduction in breakdown strength.
Therefore, it is difcult to obtain a full set of DeE loops at 1 Hz
poling frequency at 70  C.
The physical origin for the DHLs observed for P(VDF-TrFE-CFE)
terpolymers is explained in the top panel of Fig. 10. In the rst
step, pre-expansion of the interchain distance in PVDF crystals by
TrFE via repeating-unit isomorphism is important, namely, l2 > l1
where l1 and l2 are interchain distances in PVDF and P(VDF-TrFE)
crystals, respectively. This rst step expansion of interchain distance makes it possible to further incorporate the even larger third
comonomer, such as VC, CFE, CDFE, and CTFE, into the isomorphic
crystalline structure. A good example for this effect is to compare
P(VDF-CTFE) with P(VDF-TrFE-CTFE). For P(VDF-CTFE), the CTFE
repeating units are too large to be effectively incorporated in
a PVDF crystal. As a result, most CTFE units are repelled from the
PVDF crystal and no RFE behavior can be observed in P(VDF-CTFE)
[27,73]. However, CTFE can be effectively incorporated into
a P(VDF-TrFE) crystal, leading to an RFE behavior [58,87]. The
incorporation of the even larger third comonomer further expands
the interchain distance, namely, l3 > l2, where l3 is the interchain
distance of the terpolymer crystal. Note that HFP cannot be effectively incorporated into P(VDF-TrFE) to form the RFE phase because
the HFP unit is too large [88]. Meanwhile, these larger comonomers
can effectively pin the P(VDF-TrFE) chains, if they are randomly
distributed along the chain. For example, in the P(VDF-TrFE-CFE)
59.2/33.6/7.2 terpolymer, on average two neighboring CFE units
can pin about 14 VDF/TrFE repeating units. This pinning will allow
the in-between P(VDF-TrFE) chains to rotate more freely because of
the increased interchain distance. However, these CFE units can

Fig. 10. Schematic representations of physical (temporary) and chemical (permanent) pinning in laterally expanded P(VDF-TrFE) crystals, namely, P(VDF-TrFE)-based terpolymers
and e-beam crosslinked P(VDF-TrFE) copolymers. l1, l2, and l3 are average interchain distances in PVDF, P(VDF-TrFE), and P(VDF-TrFE)-based terpolymers or e-beamed P(VDF-TrFE),
respectively. The orange-red and blue ovals with arrows in them represent dipolar TrFE and the third comonomer with larger sizes. The green bars represent the chemical crosslinks
inside the crystal.

1718

L. Yang et al. / Polymer 54 (2013) 1709e1728

only provide temporary physical pinning. When the poling electrical eld becomes high enough, dipoles start to switch because
the CFE unit also has a dipole moment. Consequently, the high
electric eld will generate FE domains within the RFE matrix.
However, these FE domains should be weaker than those formed by
mechanical stretching because they can transform back to the RFE
phase after a decrease in the poling eld. We speculate that the
electric eld-induced FE domains in the terpolymer may be just
large enough so that signicant hysteresis is observed during
RFE4FE transitions.
2.3. Narrow single hysteresis loop behavior in e-beam irradiated
P(VDF-TrFE) copolymers
From the above discussion, it would be better to permanently
pin P(VDF-TrFE) chains in the crystal in order to avoid hysteresis
during the RFE4FE transitions. One effective way is to chemically
crosslink the P(VDF-TrFE) crystal with an optimized crosslinking
density (see the bottom panel in Fig. 10). Nonetheless, effective
ways to crosslink polymer crystals via chemical reactions have not
yet been possible because polymer crystals, being of greater density
than the amorphous phase, do not allow penetration of crosslinking agents such as molecular radicals into them. Instead, ionizing radiation, such as e-beam [16,89,90], g-ray [91,92], and proton
radiation [23,93e95], have been used to crosslink PVDF and PVDFbased copolymer crystals. Intriguingly, only P(VDF-TrFE) crystals
are found to be susceptible to ionizing radiation, and in this case the
interchain distance gradually increases with increasing irradiation
dose. PVDF and other PVDF-based copolymer crystals do not show
similar changes in the crystalline phase upon ionizing irradiation
[89,90]. This must be attributed to the special reactivity of TrFE to
ionizing irradiation. For example, upon e-beam irradiation the TrFE
units can transform into >CHeCF3 groups, as evidenced by hightemperature (170  C) solid-state 19F NMR [96,97], although the
detailed transformation mechanism is still unclear. Note that the
temperature for the solid-state NMR study has to be above the Tm of
P(VDF-TrFE) crystals. Otherwise, the 19F signals inside the crystals
cannot be obtained due to the low mobility of the crystalline nuclei.
Therefore, it is fairly difcult to determine explicitly what has
happened inside P(VDF-TrFE) crystals after ionizing radiation.
However, we consider that the P(VDF-TrFE) crystals must be
internally crosslinked. These internal chemical crosslinks not only
expand the interchain distance, but also permanently pin the
P(VDF-TrFE) chains in between. As a result, easy dipole rotation,
nanodomains and thus narrow hysteresis loops will be achieved.
In this study, a hot-pressed P(VDF-TrFE) 50/50 lm (ca. 60e
80 mm) is uniaxially stretched to a draw ratio of 400% at a draw

rate of 12 mm/min at room temperature. Afterwards, the lm is


annealed in the PE phase at 100  C for 2 h to achieve a high crystallinity of >85 wt.% (determined by WAXD and DSC), following
previous efforts to achieve single crystal-like P(VDF-TrFE) crystals
[69]. The nal lm thickness is about 20e23 mm. The uniaxially
stretched P(VDF-TrFE) 50/50 lm is then subjected to high-energy ebeam irradiation at 70  C for 20, 40, and 60 Mrad at a dose rate of ca.
1 Mrad/min. The DSC results for these e-beam irradiated P(VDFTrFE) 50/50 samples are shown in Fig. 11. After e-beam irradiation
at 20 Mrad (Fig. 11A), both TC (47  C, weak and broad) and Tm (143  C,
sharp) decrease as compared to the TC (64  C) and Tm (157  C) for the
pristine P(VDF-TrFE) 50/50 copolymer (see Fig. 3H). The reduced TC
may be attributed to reduced ferroelectric domain size after e-beam
irradiation. However, the reduction in Tm (14  C) is much larger than
that (4.1  C) for a biaxially oriented PVDF lm (12 mm thick) that has
been e-beam irradiated at 105  C for 20 Mrad (see Fig. S2 in the
Supplementary material and Ref. [90]). For the biaxially oriented
PVDF lm, the slight decrease in Tm is attributed to the reduced
lateral crystallite size but not necessarily to the crystalline lamellar
thickness decrease because the lamellar thickness (10e20 nm) is
less susceptible to e-beam irradiation than the lateral size. We
consider that e-beam irradiation causes permanent changes in the
P(VDF-TrFE) crystal, converting the TrFE units into chemical crosslinks. These structural defects substantially decrease the Tm after ebeam irradiation for P(VDF-TrFE). Upon melt-recrystallization, the
TC for the e-beam irradiated P(VDF-TrFE) 50/50 at 20 Mrad becomes
more obvious at 55  C, which is slightly increased as compared with
the TC at 47  C for the as-irradiated sample. This can be explained by
an increase in the ferroelectric domain size after exclusion of
crosslinked defects from the crystal through melt-recrystallization.
Meanwhile, the lamellar thickness may decrease due to crosslinks,
and the Tm therefore decreases to 139  C. Upon increasing the radiation dose to 40 and 60 Mrad, respectively, the TC disappears
during the rst heating (Fig. 11B and C), consistent with previous
results [17,98]. The reason for the disappearance of TC for these
irradiated samples may be attributed to the second-order phase
transition, and this will be discussed later. Meanwhile, the Tm values
continuously decrease signicantly. For example, at 40 Mrad the Tm
is 129  C (a 28  C decrease), and at 60 Mrad the Tm is 116  C (a 41  C
decrease). After melt-recrystallization, obvious but broad Curie
transitions reappear, and the TC values decrease with increasing the
e-beam dose; 45  C for the 40 Mrad dosed sample and 41  C for the
60 Mrad dosed sample.
From previous work on e-beam irradiated P(VDF-TrFE) [16,17], it
is known that the disappearance of TC after e-beam irradiation
above 40 Mrad at 70  C can be attributed to the formation of the
RFE phase. However, the structure of the RFE phase in e-beam

Fig. 11. DSC thermograms of the e-beam irradiated uniaxially stretched P(VDF-TrFE) 50/50 during the rst heating, rst cooling, and second heating cycles: (A) 20 Mrad, (B)
40 Mrad, and (C) 60 Mrad. The heating and cooling rates are 10  C/min.

L. Yang et al. / Polymer 54 (2013) 1709e1728

irradiated P(VDF-TrFE) is still not well-understood. From the above


three e-beam irradiated samples, the 20 Mrad sample still possess
an FE phase, and its electrical property is typically ferroelectric at
room temperature (see Fig. S3 in the Supplementary material).
Meanwhile, we observe that the ferroelectric properties of the 40
and 60 Mrad samples are quite similar (see Fig. S4 in the Supporting
material for the ferroelectric properties of the 40 Mrad sample).
Therefore, we use the 60 Mrad sample for investigation by WAXD.
Fig. 12A shows the 2D WAXD pattern for the e-beam irradiated
(60 Mrad) uniaxially stretched P(VDF-TrFE) 50/50 lm. The sharp
(110/200)RFE reection is seen on the equator and the broad
(002)RFE reection is on the meridian, indicating long-range order
in the lateral direction and disorder along the chain direction
(similar to the arrangement of the rotator phases in oligo-alkanes
and polyethylene at high pressures [99]). The d-spacing of the
(110/200) reection is 4.80 
A at 38  C and it continuously increases with increasing temperature (Fig. 12C), whereas the (002)
reection keeps constant at 2.3 
A. When the temperature is at 123e
A. During the rst
128  C, the (110/200) d-spacing increases to 5.0 

1719

heating process, again there is no obvious rst-order transition


observed in the WAXD proles in Fig. 12C (also see Fig. S1B in the
Supplementary material). After heating to 150  C, the (110/200)
reection on the equator completely disappears, leaving an isotropic amorphous halo centered at 5.3 
A. Intriguingly, the (002)
reection on the meridian does not disappear, but transforms into
a weak halo centered at 2.3 
A (see Fig. S5 in the Supplementary
material). We consider that the chain orientation in the molten
state does not completely disappear but is somewhat locked in by
e-beam crosslinking. This result also indirectly proves that the
P(VDF-TrFE) crystals are crosslinked by e-beam irradiation. The
weak halo at 2.3 
A should represent the average spacing along the
frozen chain direction with random T and G conformations. After
melt-recrystallization, the 2D WAXD pattern is shown in Fig. 12B.
The overall crystallinity (determined by DSC) for the recrystallized
sample substantially decreases to 38 wt.% due to the crosslinking,
as compared to 60 wt.% for the as-irradiated sample. The d-spacing
A at 38  C. Meanwhile,
of the (110/200)RFE decreases to 4.73 
crystal orientation is largely preserved due to the frozen oriented

Fig. 12. 2D WAXD patterns of (A) as-irradiated (60 Mrad) and (B) melt-recrystallized P(VDF-TrFE) 50/50 lm irradiated at 60 Mrad. Before irradiation, the lm was obtained by
uniaxially stretching to a 400% draw ratio followed by annealing at 100  C for 2 h. The bottom panel shows 1D WAXD proles for the equator and meridian reections during (C) the
rst and (D) the second heating cycles. The heating rate is 10  C/min and each curve is 5  C apart with the last temperature being 150  C. The inner two weak/diffuse rings belong to
the Kapton tape used to encapsulate the sample.

1720

L. Yang et al. / Polymer 54 (2013) 1709e1728

chains as mentioned above. Upon the second heating, complete


melting of the crystals is observed above 130  C, where the (110/
200) reection at 4.96 
A disappears on the equator and the (002)
reection on the meridian transforms into a weak halo at 2.3 
A.
Again, no obvious rst-order transition is observed from the WAXD
proles in Fig. 12D (see Fig. S1B in the Supplementary material). We
conclude that WAXD is not sensitive enough to detect details of
crystal structure and chain conformation in the RFE phase.
We then utilize FTIR to study the chain conformation in the RFE
phase of the e-beam irradiated (60 Mrad) uniaxially stretched
P(VDF-TrFE) 50/50 lm, and the results are shown in Fig. 13 for the
rst and second heating cycles. At 40  C in the rst heating cycle
(Fig. 13A), the absorption bands for long T sequences are seen at
475, 510, and 848 cm1, and the bands for short T and TG sequences
are observed at 526 and 615 cm1, respectively. The coexistence of
both long and short T sequences indicates that the RFE phase has
a disordered crystalline structure. However, the peak intensity of
the 510 cm1 band is stronger than that of the 615 cm1 band,
indicating that the ferroelectricity dominates in the sample
at 40  C. Upon increasing temperature, the bands at 475, 510, and
848 cm1 for the long T sequence gradually decrease in intensity,
whereas the band at 526 cm1 for the short T sequence gradually
grows in intensity. Above 50e60  C, the band at 526 cm1 becomes
stronger than that at 510 cm1. Finally, at 120  C all the crystals
melt; however, there are no noticeable changes in the FTIR spectrum as compared to the spectra just before crystal melting. This
occurs because before melting, the crystals are so defective that the
chains have already predominantly adopted the TG conformation,
which is not much different from that in the melt. After meltrecrystallization and cooling to 40  C (Fig. 13B), the long T
sequence band at 510 cm1 becomes much sharper and the short T
sequence band at 526 cm1 becomes weak. This is consistent with
the DSC results that more ferroelectricity is developed with the TC
at 41  C after melt-recrystallization for the irradiated sample (see
Fig. 11C). Upon increasing the temperature to 45  C, the short T
sequence band at 526 cm1 starts to grow, corresponding well to
the weak Curie transition at 41  C in DSC (Fig. 11C). Finally, at 120  C
all the crystals melt, but no signicant change in the FTIR spectra is
seen before and after the crystal melting. From these FTIR results,
we see that all changes are quite subtle because the RFE phase
already contains a disordered crystal structure with a signicant
amount of short T and TG sequences. Melt-recrystallization will
largely expel the structural defects such as crosslinks from the
crystals, resulting in a signicant decrease in the overall crystallinity and enhanced ferroelectricity (possibly due to enlarged ferroelectric domains). Comparing the e-beam irradiated (60 Mrad)

P(VDF-TrFE) 50/50 copolymer with the P(VDF-TrFE-CTE) 59.2/33.6/


7.2 terpolymer, we see no bands for the T3 sequence at 772 and
800 cm1. We speculate that the molar content of the structural
defects (crosslinks) may be lower than that (7.2 mol%) of CFE in the
terpolymer. As a result, long T sequences rather than the T3
sequence are adopted for the e-beam irradiated P(VDF-TrFE).
Nonetheless, this needs further investigation.
The dynamic dielectric behavior of the e-beam irradiated
(40 Mrad) uniaxially stretched P(VDF-TrFE) 50/50 lm during the
rst and second heating cycles are studied by BDS. Note that the
uniaxially stretched lm with 40 Mrad irradiation exhibits similar
ferroelectric properties to the lm with 60 Mrad irradiation.
Therefore, we use the 40 Mrad lm as an example here. As shown
in Fig. 14, both the glass transition and Curie transition for the
amorphous and crystalline dipoles are seen during heating for
r 0 and r 00 as a function of temperature. For example, during the
rst heating cycle, the Tg and TC are observed at 10 and 31  C
(1 Hz) in the r 00 plot, and they gradually increase with increasing
frequency (Fig. 14B). During the second heating cycle after meltrecrystallization, the Tg and TC are seen at 20 and 40  C (1 Hz)
in the r 00 plot, and both values increase with increasing frequency
(Fig. 14D). Before melt-recrystallization, the RFE/PE Curie transitions under various frequencies appear to be broad in the r 0
plot (Fig. 14A) and weak in the r 00 plot (Fig. 14B) for the asirradiated lm. After melt-recrystallization, the FE/PE Curie
transitions become sharper in the r 0 plot (Fig. 14C) and much
more pronounced in the r 00 plot (Fig. 14D). Similar results have
been reported in the literature [100,101]. This clearly indicates
that certain ferroelectricity is regained for the melt-recrystallized
lm due to the exclusion of structural defects (mostly likely
chemical crosslinks) outside the P(VDF-TrFE) crystals. The hightemperature (>50  C) and low-frequency (<1 kHz) upturns in
the r 00 in Fig. 14B and D again are attributed to the impurity ion
relaxation, as revealed by the frequency scans for the rst and
second heating cycles (Fig. 14EeH). In Fig. 14E and G, upturns in
the r 0 are observed at low frequencies (<1 Hz) and high temperatures (>75  C). Meanwhile, in Fig. 14F and H upturns in the
r 00 in the double-log plot show a slope of ca. 1, clearly indicating
the migrational loss from impurity ions [80,85]. At high frequencies, a decrease in r 0 and a broad peak in r 00 can be
attributed to the relaxation of amorphous dipole. At intermediate
frequencies, plateaus in r 0 are observed in Fig. 14E and G. For
both the rst and second heating cycles, the maximum plateau r 0
values are observed at 50  C, because the TCs are fairly close
(around 31e40  C) for the as-irradiated and melt-crystallized
lms.

Fig. 13. FTIR spectra for the e-beam irradiated (60 Mrad) uniaxially stretched P(VDF-TrFE) 50/50 lm during (A) the rst and (B) the second heating (after annealing at 115  C for
10 min) cycles. The spectra are collected during a stepwise heating process from 40 to 120  C with each curve 5  C apart. All curves are overlaid for clarity.

L. Yang et al. / Polymer 54 (2013) 1709e1728

1721

Fig. 14. (A and C) Real r 0 and (B and D) imaginary r 00 relative permittivity as a function of temperature under different frequencies for the e-beam irradiated (40 Mrad)
uniaxially stretched P(VDF-TrFE) 50/50 lm during (A and B) the rst and (C and D) the second heating (i.e., after melt-recrystallization) cycles. (E and G) r 0 and (F and H) r 00 as
a function of frequency at different temperatures for the e-beam irradiated (40 Mrad) uniaxially stretched P(VDF-TrFE) 50/50 lm during (E and F) the rst and (G and H) the second
stepwise heating (i.e., after melt-recrystallization) cycles.

Temperature-dependent ferroelectric properties of the e-beam


irradiated (60 Mrad) uniaxially stretched P(VDF-TrFE) 50/50 lm
are studied by DeE loop measurements, and the results are shown
in Fig. 15. At 25  C, the DeE loops are relatively open, and this can
be attributed to the slightly increased domain size at such a low
temperature. On increasing temperature, the hysteresis in DeE
loops gradually decreases and the most narrow hysteresis loops
are observed at 25  C. When the temperature increases to 50  C, the
DeE loops start to open up again, especially at 1 Hz. Finally, the
hysteresis loops substantially open up for 1 Hz at 75  C. The opened
loops at high temperatures and low frequencies can be attributed to
enhanced ion migrational loss and/or dc conduction in the sample
[80,85]. For all these temperatures, the Dmax slightly increases with
decreasing poling frequency; however, the loop shape does not
change much except for 1 Hz at 75  C.
This temperature-dependent RFE behavior can be explained
using the schematic representation in the bottom panel of Fig. 10.
We hypothesize that the inclusion of TrFE is critical for e-beam
induced crosslinking inside the crystals, because PVDF and its
other copolymers do not show the RFE behavior even after e-beam
irradiation. After high-energy e-beam irradiation, the crystal is
internally crosslinked and this is important for the appearance of
the RFE behavior. When the crosslinking density is low (e.g.,
20 Mrad at 70  C), the sample is still ferroelectric with a decreased
TC (see Fig. 11A). When the crosslinking density increases to
a critical value (e.g., with 40 and 60 Mrad irradiation at 70  C), the
true RFE behavior can be achieved without any observation of TC
(see DSC curves during the rst heating in Fig. 11B and C). We
consider that the Curie transition has become a second-order
transition. The crosslinking of both amorphous and crystalline
phases has prevented detailed structural analyses of what has
happened in the crystal. However, the crosslinks inside the crystal
permanently pin P(VDF-TrFE) chains with suitable segmental
lengths, resulting in nanodomains (see the bottom panel of
Fig. 10). Therefore, the pinned P(VDF-TrFE) chains in between
behave like torsional springs and the conventional ferroelectricto-paraelectric Curie transition is effectively prohibited. Meanwhile, the interchain distance is also enlarged, as evidenced by the
WAXD study in Fig. 12. On applying an external electric eld, the

dipoles in the pinned P(VDF-TrFE) chains can rotate more freely.


Due to the tendency of permanent pinning and enlarged interchain distance to reduce the dipoleedipole interaction, the ferroelectric domains must be small in both directions perpendicular
and parallel to the chains. It is these nanodomains with easy
rotating main-chain dipoles that result in a true RFE behavior (i.e.,
narrow single hysteresis loops) in e-beam irradiated P(VDF-TrFE).
Due to the lack of effective experimental methods to determine
the size and shape of ferroelectric domains at the nanoscale, we
resort to computer simulation to visualize these nanodomains
later. Finally, when the crosslinking density is too high in the
crystal, P(VDF-TrFE) chains will be permanently locked in and the
dipole switching will be lost as observed in previous reports
[16,17]. Although the chemical pinning effectively eliminated the
TC in the sample, the size of these nanodomains may still be
temperature-dependent. At low temperatures (e.g., 25  C), the
nanodomains appear larger and thus the DeE loops become broad
(see Fig. 15AeD). As the size of nanodomains gradually decreases,
narrow hysteresis loops are observed (e.g., at 0e50  C). On further
increasing the temperature, the mobility of P(VDF-TrFE) chains
becomes so high that both impurity ion loss and dc conduction
signicantly increase, as discussed above. As a result, the DeE
loops will eventually broaden at high temperatures and low
frequencies.
After melt-recrystallization of the e-beam irradiated P(VDF-TrFE)
lms, structural defects (i.e., the pinning points) will be largely
excluded from the crystals. As a result, certain ferroelectricity is
restored, as proven by the reappearance of TC during the second
heating cycles in DSC (see Fig. 11B and C). The reappearance of
ferroelectricity after melt-recrystallization results in broader DeE
hysteresis loops, as shown in Fig. 16 [101]. Before melt-recrystallization, typical RFE behavior is observed for the e-beam irradiated uniaxially stretched P(VDF-TrFE) 65/35 lms with different
doses; 60, 85, and 100 Mrad (1.2 MeV e-beam irradiation at 100  C).
After melt-recrystallization, the DeE loops broaden with an
increased Dmax. The lower Dmax for the e-beam irradiated P(VDFTrFE) lms suggests fewer polarizable VDF/TrFE dipoles and thus
smaller domains due to the chemical pinning (or crosslinks) inside
the crystals.

1722

L. Yang et al. / Polymer 54 (2013) 1709e1728

Fig. 15. Bipolar hysteresis loops for the e-beam irradiated (60 Mrad) uniaxially stretched P(VDF-TrFE) 50/50 lm (draw ratio 400%) at different temperatures: (AeD) 25  C, (EeH) 0  C,
(IeL) 25  C, (MeP) 50  C, and (QeT) 75  C. The poling frequencies decrease from left to right: (A, E, I, M, Q) 1000 Hz, (B, F, J, N, R) 100 Hz, (C, G, K, O, S) 10 Hz, and (D, H, L, P, T) 1 Hz,
respectively. All poling eld (a triangular waveform) starts at 25 MV/m with step increments of 25 MV/m except those at 1000 Hz.

2.4. Computer simulation of nanodomains in pinned PVDF chains


As we mentioned above, it is not possible to visualize ferroelectric domains, especially nanodomains, using current experimental techniques. To understand the nature of nanodomains and
their relationship to RFE behavior, we have performed fully atomistic simulations of pinned zig-zag PVDF chains in a crystal with
enlarged lateral unit cell dimensions (i.e., ax and by). In this molecular dynamics simulation, all the two-body, three-body, and
four-body interaction constants depend on the type of atoms

involved in the interaction. These parameters, as well as the van der


Waals interaction parameters and the atomic partial charges, are
taken from the force-eld developed for PVDF polymers [102]. The
long-range electrostatic interactions between charged particles are
handled using the standard Lekner summation algorithm [103].
The initial conguration is a zero-temperature crystal with all trans
chains running parallel to each other along the z-axis (Fig. 17A).
Note that the unit cell c-axis is along the chain direction or the zaxis. Each chain contains 30 e(CH2CF2)e repeating units. There are
in total 144 such chains, corresponding to 72 PVDF ab-cells in the xy

L. Yang et al. / Polymer 54 (2013) 1709e1728


100

100

100

C
2

50

P (mC/m )

50

P (mC/m )

50

P (mC/m )

1723

-50

-50

-50
100 Mrad

85 Mrad

60 Mrad
R e c ry s ta lliz e d
-100
-1 0 0

-5 0

50

100

150

R e c rys ta lliz e d

Recrystallized
-100

-1 5 0

-100
-1 5 0

-1 0 0

-5 0

E (MV/m)

E (MV/m)

50

100

150

-1 5 0

-1 0 0

-5 0

50

100

150

E (MV/m)

Fig. 16. Bipolar hysteresis loops for e-beam irradiated uniaxially stretched P(VDF-TrFE) 65/35 lms before (blue) and after (red) melt-recrystallization: (A) 60 Mrad, (B) 85 Mrad, and
(C) 100 Mrad. The poling frequency is 10 Hz with a triangular waveform. (Reproduced and adapted with permission from Ref. [101]. Copyright 2005 American Institute of Physics.)

plane. The end CH2 and CF2 groups are elastically attached to the
surfaces of two parallel walls separated by a distance, D 30cz. On
the surface of the left wall placed at z 0 nm, all the xed CH2
groups are oriented in such a way that their H atoms point in the
negative direction of the x-axis. Correspondingly, the F atoms of the
xed CF2 groups on the surface of the right wall, placed at z D,
point along the positive direction of the x-axis. Such a rm xing of
both chain ends mimics the chemical crosslinks in the e-beam
irradiated P(VDF-TrFE) crystal and prevents their rotation around
the z-axis near the walls. During the initial stage of the simulation,
the system is gradually raised from zero to room temperature in 6
consecutive runs at increments of 50 K. In each run the system is
equilibrated for 20 ps. Once the system temperature reaches 300 K,
simulation data, such as the positions of atoms, their velocities, and

the forces acting on them, are stored at time intervals of 1 ps for


processing later.
Although the lattice constants of the pure b-phase unit cell are
ax 0.85 nm, by 0.491 nm, and cz 0.256 nm, other values with
enlarged by, but with a xed by/ax ratio are also considered in our
simulations. We nd that increasing the separation distance between chains, while keeping the geometry of the crystal intact,
strongly affects the average dipole moment of the sample and reduces the strength of Coulomb correlations between aligned dipoles. When the separation distance becomes greater than the
range of the van der Waals interaction, the chains no longer feel the
steric hindrance of their neighbors. As a consequence, the dipoles in
each chain can respond more readily to temperature uctuations.
Our simulation results (not shown here) indicate that at smaller

Fig. 17. Simulation results for the time evolution of a layered domain morphology in a PVDF crystal with all trans chains along the z-axis sandwiched between two parallel walls.
The unit cell of the PVDF crystal is enlarged about four times in both x and y directions to ax 3.5 nm and by 2.0 nm. (AeJ) show the morphology at simulation times of 1, 2, 3, 4, 5,
10, 50, 100, 200, and 250 ps. The domains with dipoles in the upward direction are shown in colors from blue to red, corresponding to different dipole amplitudes. Blue color
corresponds to the dipole amplitude of 0 D whereas red color corresponds to 2.1 D. Empty regions represent the domains with their dipole moments down (or negative amplitude).
(KeO) show domain structures for the middle layer at a simulation time of 300 ps. Successive images show dipole strength varying from 0 to 2 D with 0.5 D increments. All
simulation boxes have dimensions of 21.0  24.0  7.68 nm3.

1724

L. Yang et al. / Polymer 54 (2013) 1709e1728

separations, i.e., by < 1 nm, the average dipole moment of a system


having chains with 100 repeating units decreases only slightly from
its zero-temperature value of 2.1 D to 1.8 D. At larger separations,
the average dipole moment drops sharply, and the uniform dipole
orientation found in the b-form unit cell is absent in regions far
away from both walls.
In a preliminary simulation, we set by at 0.6, 1.2, and 2.0 nm,
respectively, while keeping the ratio of by/ax xed as in the b-form.
For by 0.6 nm, no dipole rotation is observed for simulation times
up to 500 ps. For by 1.2 nm, strong chain aggregation leads to
a non-uniform density distribution in space. When by 2 nm, no
chain aggregation is observed and a unique nanodomain structure
evolves from a single well-aligned domain (see Fig. 17AeJ). Within
10 ps, tiny domains with different dipole orientations start to
appear and continue to grow in size with increasing time. Beyond
50 ps, it is found that an isolated middle section of the simulated
system keeps the dipole moment oriented upward, while the regions between the walls and the middle section change their
dipolar orientation from the upward direction to the downward
direction. Up to 300 ps, the nanodomain morphology with four
layers (note that the left and right layers at the wall are considered
as one full layer) seems to become stable. We consider that the
layered domain structure must be dictated by the two conning
walls. Each nanodomain layer has a length of about 7e8 repeating
units. The lateral dimension of these layered nanodomains is portrayed in Fig. 17KeO by showing the dipole amplitudes in the
middle, upward-oriented region at a simulation time of 300 ps.
Upon increasing the threshold amplitude of the displayed dipole
moment from 0 to 2 D in 0.5 D increments, the lateral dimension of
the nanodomains gradually decrease from bicontinuous to more
isolated. If we use an amplitude of 1.5 D (Fig. 17N) for the domain
size analysis, the average lateral dimension is about 2e4 unit cells.
It is noticed that the chain separation distances employed in the
simulation are set at unrealistically large in order to achieve the
development of nanodomains. The necessity to consider such extremes arises from the limitations encountered in performing fully
atomistic simulations. With simulated times of even 500 ps,
insufcient molecular realignment would be observed at smaller
separations. Consequently, we are obliged to compensate by
speeding up the relaxation processes articially. If unlimited computer time were available, one might expect to see, in systems with
only a 12% expansion of the interchain distance, the type of
behavior we nd at a 400% increase in separation. We are currently

exploring other avenues by which the simulation may be accelerated in order to obtain more realistic results.
3. Effect of crystaleamorphous interaction on novel
ferroelectric behaviors (the nanoconnement effect)
As mentioned above, most ferroelectric polymers are semicrystalline (we have not focused on ferroelectric liquid crystal
polymers [104] in this article). In addition to the more important
effect of the internal crystal structure, the interactions that occur at
the interface between the crystalline and amorphous regions can
also affect the ferroelectric properties of polymers. We call this the
nanoconnement effect. In a recent perspective article [13], we
have already discussed this nanoconnement effect and here we
will briey review our conclusions.
In a simple case, a semicrystalline ferroelectric polymer can be
represented by the serial capacitor model shown in Fig. 18A. In this
model, a ferroelectric polymer (e.g., PVDF) crystal is sandwiched
between two amorphous layers, which are in direct contact with
two metal electrodes. The chain direction is perpendicular to the
external eld. Upon application of an external electric eld (E0)
above the coercive eld (Ec), dipoles in the ferroelectric PVDF
crystal will align along the eld direction, creating a local depolarization eld (Edepol) in the crystal, as shown in Fig. 18A [105,106]:

Edepol

Pin
0

(1)

where Pin is the polarization inside the ferroelectric crystal due to


the electrically aligned dipoles. Meanwhile, a compensation polarization (Pcomp) in the amorphous PVDF at the crystale
amorphous interface will form outside the ferroelectric crystal.
Note that Pin may not equal Pcomp, especially during a dynamic
poling process. Based on the serial capacitor model in Fig. 18A, the
local polarization eld for the ferroelectric crystal is imposed from
the charges on the electrodes and can be dened as [105,106]:

Epol

Q Pcomp
D

0
0

(2)

where Q is the charge density in a vacuum capacitor, i.e., Q 0E0.


Dipole switching depends on the local electric eld in the crystal
(EL,cryst), which has at least three sources of contribution; i) Epol

Fig. 18. Schematic illustrations of (A) an electrically polarized ferroelectric PVDF crystal (blue rectangles) sandwiched between two amorphous layers and (B) a nanoconned
ferroelectric P(VDF-TrFE) crystal by a thin layer of polystyrene (PS, in green color) under a poling electric eld. The folded chain direction in the lamellar crystals is perpendicular to
the dipole moment or the external electric eld E0. Both polarization and depolarization elds are shown in the schemes. (Reproduced with permission from Ref. [72]. Copyright
2011 American Chemical Society.)

L. Yang et al. / Polymer 54 (2013) 1709e1728

(positive) ii) Edepol (negative) and iii) other contributions from dipoles nearby [105,107]. Note that there should not be contribution
from both amorphous layers to the local eld inside the crystal. If
we consider that contribution from nearby dipoles in the crystal
can cancel out [105,107], we may assume that the contribution (iii)
can be neglected. Then, the local electric eld largely depends on
the competition between Epol (or Q Pcomp) and Edepol (or Pin). Upon
a forward poling (to the positive direction), Q Pcomp is larger than
Pin and thus Epol will be higher than Edepol. Crystalline dipoles will
be polarized along the external electric eld direction. Upon
reverse poling (to the negative direction), two scenarios can take
place. First, if Q Pcomp is larger than Pin as both Edepol and Epol
decrease with the external eld, the Epol will be higher than the
Edepol and the normal ferroelectric behavior with a single poling
process will be resulted. Second, if Pin is higher than Q Pcomp, then
Edepol will be greater than Epol at a certain point. As a result, fully
aligned dipoles/domains will become less stable than they are in
a situation where part of mobile (or reversible) dipoles/domains
reverts to the opposite direction, resulting in a minimum (or even
net zero) remanent polarization. Therefore, the DHL behavior with
a two-step poling process will be obtained [38,72]. Finally, in
a recent report we observe that there are a certain amount of
reversible dipoles/domains in P(VDF-HFP) (and also PVDF), in
addition to irreversible ones [108]. These reversible (or mobile)
dipoles/domains can relax (i.e., randomize) within a few ms upon
removing the poling electric eld. As a result, the same discharged
energy density will be obtained regardless of how fast the sample is
charged and discharged up to 1000 Hz. From the above discussion
of the internal crystal structure effect, we consider that even in
close-packed PVDF and its copolymer crystals, a certain number of
nanodomains may still exist depending on the crystallite size and
poling condition. These nanodomains should be fairly mobile or
reversible because of small spontaneous polarization. If the amount
of these reversible nanodomains is high enough (e.g., close to 50%),
a minimum remanent polarization will be achieved around E 0,
resulting in a two-step poling process or the DHL behavior.
On the basis of the above discussion, the DHL behavior can be
achieved by either increasing the Pin or decreasing the Pcomp, or
a combination of both. Generally speaking, it is relatively difcult to

1725

increase the Pin, because the dipole moment per repeating unit is
xed by the covalent bonds. On the contrary, reducing the Pcomp is
easier and can be effectively used to obtain the DHL behavior. For
example, the average dipole moment per repeating unit for the
amorphous P(VDF-TrFE) is lower than that for the amorphous
PVDF. The Pcomp for P(VDF-TrFE) therefore should be lower than
that for PVDF. This explains why the DHL behavior is often observed
for P(VDF-TrFE) (especially with relatively high TrFE contents)
rather than PVDF at certain poling electric elds [5,109e115].
Nonetheless, the DHL behavior in P(VDF-TrFE) is not stable. After
repeated electric poling (e.g., 5e10 times), the double hysteresis
eventually disappears and is replaced by the single normal ferroelectric loop [111,113]. This can be explained by the growth of
nanodomains into large domains and thus an increase in Pcomp
upon repeated electric poling in P(VDF-TrFE) samples.
To achieve stable DHLs, we propose to conne PVDF or P(VDFTrFE) crystals with a thin layer of low polarizability polymer such
as polystyrene (PS) to purposely reduce Pcomp and limit domain
growth upon repeated poling (see Fig. 19B) [13,72]. In detail, PS side
chains are grafted from a P(VDF-TrFE-CTFE) 88.2/7.7/3.5 terpolymer
using atom transfer radical polymerization (ATRP) [116], following
an established method in literature (see the top panel of Fig. 19)
[117]. The last step dechlorination by n-Bu3SnH is conducted to
avoid crosslinking of the graft copolymer under further thermal
treatments. The PS content is determined to be 14 wt.%, corresponding to 3.2 styrene repeating units per PS graft and each mainchain contains 47 PS side chain. A P(VDF-TrFE) 93/7 (mol/mol)
copolymer is used as a control to compare with the graft copolymer.
P(VDF-TrFE) 93/7 and P(VDF-TrFE-CTFE)-g-PS(14%) lms are
obtained by hot-pressing at 240  C followed by uniaxial stretching
at 110  C. The 2D XRD patterns in the insets of Fig. 19A and B
indicate that the chains in the P(VDF-TrFE) crystals are parallel to
the stretching direction. Bipolar DeE hysteresis loops for P(VDFTrFE) 93/7 and P(VDF-TrFE-CTFE)-g-PS(14%) are shown in Fig. 19A
and B, respectively, and a clear difference is seen. For P(VDF-TrFE)
93/7, DHLs are seen at the poling eld below 200 MV/m, above
which only normal ferroelectric loops are seen. Note that these low
eld DHLs are unstable and will disappear after repeated poling as
we mentioned above. For P(VDF-TrFE-CTFE)-g-PS(14%), DHLs are

Fig. 19. (Top) Synthesis of P(VDF-TrFE-CTFE)-g-PS graft copolymer from the P(VDF-TrFE-CTFE) terpolymers using atom transfer radical polymerization followed by dechlorination.
(A) and (B) Bipolar DeE hysteresis loops for uniaxially stretched P(VDF-TrFE) 93/7 and P(VDF-TrFE-CTFE)-g-PS(14%) lms. Corresponding 2D XRD patterns are shown as insets. (C)
The DeE loop for the graft copolymer at a poling eld of 150 MV/m. Two-step poling processes are shown to explain the competition between the Edepol and Epol. (Reprinted with
permission from Ref. [72]. Copyright 2011 American Chemical Society.)

1726

L. Yang et al. / Polymer 54 (2013) 1709e1728

seen at the poling eld up to 400 MV/m. These DHLs are stable and
can survive hundreds of poling cycles (data not shown).
The observed DHL behavior in P(VDF-TrFE-CTFE)-g-PS (14%)
can be explained using the model in Figs. 18B and 19C. After the
crystallization-induced microphase separation, a low polarizability
PS-rich interfacial layer connes (or insulates) the ferroelectric
P(VDF-TrFE) crystal. As a result, the Pcomp in P(VDF-TrFE-CTFE)-gPS(14%) greatly decreases, as compared with that in P(VDF-TrFE)
(nearly 40% at 400 MV/m). A typical DHL with the remanent
electric displacement, Drem w 0, at a poling eld of 150 MV/m is
shown in Fig. 19C. Two steps during the reverse poling process are
observed. During the rst step, Q Pcomp < Pin and thus
Epol < Edepol. The reversible dipoles/domains will revert back to the
opposite direction, resulting in nearly zero net dipole moment.
During the second step, Q Pcomp > Pin and thus Epol > Edepol. The
dipoles/domains will be aligned along the reverse eld direction.
Due to the nanoconnement effect, no growth of reversible
nanodomains is allowed. Therefore, the DHL behavior becomes
stable.

This has been seen in a P(VDF-TrFE-CTFE)-g-PS(14%) graft


copolymer.
What we have learned in this work will stimulate several new
research directions. First, the consequences of nanodomain and
pinning should be applicable to other ferroelectric polymers, such
as odd-numbered nylons, polyurethanes, and polyureas. The key is
to achieve repeating-unit isomorphism and pinning of the ferroelectric crystal. Second, other in-situ crystal crosslinking mechanisms, which will not involve ionizing radiation, should be pursued.
For example, UV-induced topochemical polymerization has been
used to achieve solid-state crosslinking of diacetylene crystals
[118,119]. If this concept can be employed for ferroelectric polymer
crystals, UV-crosslinking will be much more appealing for practical
applications. Third, a combination of both crystal internal structure
and nanoconnement effects will be able to produce double loops
with minimum hysteresis. This is more desirable for high-energy
density/low-loss electric energy storage [13]. All these opportunities make ferroelectric polymers attractive for new and better
applications.

4. Concluding remarks and outlook

Acknowledgments

Ferroelectrics and dielectrics represent two extremes of electrical properties for insulating materials. By gradually decreasing
the nonlinearity, novel ferroelectric behaviors such as narrow single (or the RFE) and DHL behaviors can be achieved. The key is to
introduce one or more permanent dipoles in a small ferroelectric
domain (or nanodomain) with enhanced dipole reversibility, surrounded by a weakly interacting matrix. In other words, nanodomains can decrease the spontaneous (or remanent) polarization,
and then enhanced dipolar reversibility can reduce polarization
hysteresis.
From the above studies of P(VDF-TrFE)-based terpolymers and
e-beam irradiated copolymers, we learn that these novel ferroelectric properties can be successfully achieved via careful tailoring
of both the crystal internal structure and the crystaleamorphous
interaction (the nanoconnement effect) in semicrystalline polymers (note that dipolar glass and liquid crystal polymers are not
considered in this feature article). In general, the effect of crystal
internal structure is more important than the nanoconnement
effect, and can be effectively achieved by repeating-unit isomorphism and by pinning inside the crystal. The effect of
repeating-unit isomorphism is to increase the interchain distance
for dipole ipping with reduced friction in the crystal, whereas
pinning will limit the domain size and provide the pinning force for
fast dipole reversal. Pinning can be either physical (i.e., temporary)
or chemical (i.e., permanent). For physical pinning, the RFE phase is
stable at low electric elds and will transform into an FE phase at
high electric elds. Reversible RFE4FE phase transitions as
a function of electric eld result in a DHL behavior. This is exactly
what is observed for P(VDF-TrFE)-based terpolymers. For chemical
pinning, the isomorphic crystal needs to be crosslinked in the solidstate (or in-situ) to achieve the permanent pinning force for
reversible dipole switching. Nonetheless, in-situ crosslinking of
ferroelectric polymer crystals has been difcult to achieve except
by the use of ionization radiations such as high-energy e-beam, gray, and proton irradiations. Once the material is properly crosslinked with an appropriate length in between neighboring crosslinking points, nanodomain and dipole reversibility (i.e., the RFE
behavior) can be obtained. This is observed for e-beam irradiated
P(VDF-TrFE) crystals. The nanoconnement effect, in general, is less
important for novel ferroelectric behaviors in semicrystalline ferroelectric polymers. Nonetheless, by manipulation of the crystale
amorphous interface to reduce the Pcomp, a two-step dipole reversal (i.e., DHL) behavior is observed due to a stronger Edepol than Epol.

LZ and QMZ thank the Army Research Ofce (ARO, W911NF-111-0534) for supporting this work. LZ also acknowledges support
from the Ofce of Naval Research (ONR, N00014-05-1-0338) and
National Science Foundation (NSF, DMR-0907580). PLT acknowledges support from the Petroleum Research Fund of the American
Chemical Society (PRF# 51995-ND7). The authors are indebted to
Professors Donald Schuele and Jerome Lando at Case Western
Reserve University and Professor Takeo Furukawa at Tokyo University of Science for helpful discussions.
Appendix A. Supplementary data
Supplementary data related to this article can be found at http://
dx.doi.org/10.1016/j.polymer.2013.01.035.
References
[1] Ducharme S, Reece TJ, Othon CM, Rannow RK. IEEE Transactions on Device
and Materials Reliability 2005;5(4):720e35.
[2] Ling QD, Liaw DJ, Zhu C, Chan DSH, Kang ET, Neoh KG. Progress in Polymer
Science 2008;33(10):917e78.
[3] Hu Z, Tian M, Nysten B, Jonas AM. Nature Materials 2009;8(1):62e7.
[4] Naber RCG, Asadi K, Blom PWM, de Leeuw DM, de Boer B. Advanced Materials 2010;22(9):933e45.
[5] Furukawa T, Takahashi Y, Nakajima T. Current Applied Physics 2010;10(1):
E62e7.
[6] Kang SJ, Bae I, Shin YJ, Park YJ, Huh J, Park SM, et al. Nano Letters 2011;11(1):
138e44.
[7] Nalwa HS. Ferroelectric polymers: chemistry, physics, and applications. New
York: Marcel Dekker, Inc.; 1995.
[8] Chen QX, Payne PA. Measurement Science and Technology 1995;6(3):249e67.
[9] Zhang QM, Li H, Poh M, Xia F, Cheng ZY, Xu H, et al. Nature 2002;419(6904):
284e7.
[10] Liu Y, Tian G, Wang Y, Lin J, Zhang Q, Hofmann HF. Journal of Intelligent
Material Systems and Structures 2009;20(5):575e85.
[11] Brochu P, Pei Q. Macromolecular Rapid Communications 2010;31(1):10e36.
[12] Chang C, Tran VH, Wang J, Fuh YK, Lin L. Nano Letters 2010;10(2):726e31.
[13] Zhu L, Wang Q. Macromolecules 2012;45(7):2937e54.
[14] Chiang Y-M, Birnie DP, Kingery WD. Physical ceramics: principles for ceramic
science and engineering. New York: J. Wiley; 1997.
[15] Furukawa T, Seo N. Japanese Journal of Applied Physics, Part 1: Regular Papers Short Notes and Review Papers 1990;29(4):675e80.
[16] Zhang QM, Bharti V, Zhao X. Science 1998;280(5372):2101e4.
[17] Bharti V, Zhao XZ, Zhang QM, Romotowski T, Tito F, Ting R. Materials
Research Innovations 1998;2(2):57e63.
[18] Fukada E. IEEE Transactions on Ultrasonics Ferroelectrics and Frequency
Control 2000;47(6):1277e90.
[19] Cheng ZY, Bharti V, Xu TB, Xu H, Mai T, Zhang QM. Sensors and Actuators A:
Physical 2001;90(1e2):138e47.
[20] Jayasuriya AC, Schirokauer A, Scheinbeim JI. Journal of Polymer Science, Part
B: Polymer Physics 2001;39(22):2793e9.

L. Yang et al. / Polymer 54 (2013) 1709e1728


[21] Xia F, Cheng Z, Xu H, Li H, Zhang Q, Kavarnos GJ, et al. Advanced Materials
2002;14(21):1574e7.
[22] Huang C, Klein R, Xia F, Li HF, Zhang QM, Bauer F, et al. IEEE Transactions on
Dielectrics and Electrical Insulation 2004;11(2):299e311.
[23] Guo S, Zhao XZ, Zhou Q, Chan HLW, Choy CL. Applied Physics Letters 2004;
84(17):3349e51.
[24] Bauer F, Fousson E, Zhang QM. IEEE Transactions on Dielectrics and Electrical
Insulation 2006;13(5):1149e54.
[25] Chen Q, Ren K, Chu B, Liu Y, Zhang QM, Bobnar V, et al. Ferroelectrics 2007;
354(1):178e91.
[26] Kochervinskii VV. Crystallography Reports 2009;54(7):1146e71.
[27] Chu B, Zhou X, Ren K, Neese B, Lin MR, Wang Q, et al. Science 2006;
313(5785):334e6.
[28] Chu B, Zhou X, Neese B, Zhang QM, Bauer F. Ferroelectrics 2006;331(1):35e42.
[29] Chu B, Zhou X, Neese B, Zhang QM, Bauer F. IEEE Transactions on Dielectrics
and Electrical Insulation 2006;13(5):1162e9.
[30] Zhou X, Chu B, Neese B, Lin M, Zhang QM. IEEE Transactions on Dielectrics
and Electrical Insulation 2007;14(5):1133e8.
[31] Zhang Z, Chung TCM. Macromolecules 2007;40(26):9391e7.
[32] Cai L, Wang X, Darici Y, Zhang J, Dowben PA. Journal of Chemical Physics
2007;126(12):124908.
[33] Lu Y, Claude J, Norena-Franco LE, Wang Q. Journal of Physical Chemistry B
2008;112(34):10411e6.
[34] Wang Y, Zhou X, Chen Q, Chu B, Zhang QM. IEEE Transactions on Dielectrics
and Electrical Insulation 2010;17(4):1036e42.
[35] Zhang L. Europhysics Letters 2010;91(4):47001.
[36] Li W, Meng Q, Zheng Y, Zhang Z, Xia W, Xu Z. Applied Physics Letters 2010;
96(19):192905.
[37] Guan F, Pan J, Wang J, Wang Q, Zhu L. Macromolecules 2010;43(1):
384e92.
[38] Guan F, Wang J, Pan J, Wang Q, Zhu L. Macromolecules 2010;43(16):
6739e48.
[39] Guan F, Wang J, Pan J, Wang Q, Zhu L. IEEE Transactions on Dielectrics and
Electrical Insulation 2011;18(4):1293e300.
[40] Wang Q, Zhu L. Journal of Polymer Science, Part B: Polymer Physics 2011;
49(20):1421e9.
[41] Neese B, Chu B, Lu SG, Wang Y, Furman E, Zhang QM. Science 2008;
321(5890):821e3.
[42] Neese B, Lu SG, Chu B, Zhang QM. Applied Physics Letters 2009;94(4):
042910.
[43] Lu SG, Zhang QM. Advanced Materials 2009;21(19):1983e7.
[44] Liu PF, Wang JL, Meng XJ, Yang J, Dkhil B, Chu JH. New Journal of Physics
2010;12:023035.
[45] Li X, Qian XS, Lu SG, Cheng J, Fang Z, Zhang QM. Applied Physics Letters
2011;99(5):052907.
[46] Lu SG, Rozic B, Zhang QM, Kutnjak Z, Neese B. Applied Physics Letters 2011;
98(12):122906.
[47] Rozic B, Lu SG, Kutnjak Z, Neese B, Zhang QM. Ferroelectrics 2011;422(1):
81e5.
[48] Valant M. Progress in Materials Science 2012;57(6):980e1009.
[49] Lu SG, Rozic B, Zhang QM, Kutnjak Z, Pirc R. Applied Physics A: Materials
Science & Processing 2012;107(3):559e66.
[50] Bobnar V, Li X, Casar G, Erste A, Glinsek S, Qian X, et al. Journal of Applied
Physics 2012;111(8):083515.
[51] Chen XZ, Qian XS, Li X, Lu SG, Gu HM, Lin M, et al. Applied Physics Letters
2012;100(22):222902.
[52] Pirc R, Kutnjak Z, Blinc R, Zhang QM. Journal of Applied Physics 2011;110(7):
074113.
[53] Pirc R, Rozic B, Kutnjak Z, Blinc R, Li XY, Zhang QM. Ferroelectrics 2012;
426(1):38e44.
[54] Samara GA. Journal of Physics: Condensed Matter 2003;15(9):R367e411.
[55] Bokov AA, Ye ZG. Journal of Materials Science 2006;41(1):31e52.
[56] Ohwada K, Tomita Y. Journal of the Physical Society of Japan 2010;79(1):
011012.
[57] Cheng ZY, Zhang QM, Bateman FB. Journal of Applied Physics 2002;92(11):
6749e55.
[58] Chung TC, Petchsuk A. Macromolecules 2002;35(20):7678e84.
[59] Bobnar V, Vodopivec B, Levstik A, Kosec M, Hilczer B, Zhang QM. Macromolecules 2003;36(12):4436e42.
[60] Klein RJ, Xia F, Zhang QM, Bauer F. Journal of Applied Physics 2005;97(9):
094105.
[61] Zhang S, Klein RJ, Ren K, Chu B, Zhang X, Runt J, et al. Journal of Materials
Science 2006;41(1):271e80.
[62] Zhang S, Chu B, Neese B, Ren K, Zhou X, Zhang QM. Journal of Applied Physics
2006;99(4):044107.
[63] Wang Y, Lu SG, Lanagan M, Zhang QM. IEEE Transactions on Ultrasonics
Ferroelectrics and Frequency Control 2009;56(3):444e9.
[64] Bauer F. IEEE Transactions on Dielectrics and Electrical Insulation 2010;
17(4):1106e12.
[65] Wunderlich B. Macromolecular physics, vol. 1. New York: Academic Press; 1973.
[66] Aharoni SM. n-Nylons: their synthesis, structure, and properties. New York:
J. Wiley & Sons; 1997.
[67] Tashiro K. Crystal structure and phase transition of PVDF and related
copolymers [Chapter 2]. In: Nalwa HS, editor. Ferroelectric polymers: chemistry, physics, and applications. New York: Marcel Dekker, Inc.; 1995. p. 63e182.

1727

[68] Pan P, Inoue Y. Progress in Polymer Science 2009;34(7):605e40.


[69] Ohigashi H, Omote K, Gomyo T. Applied Physics Letters 1995;66(24):
3281e3.
[70] Omote K, Ohigashi H, Koga K. Journal of Applied Physics 1997;81(6):2760e9.
[71] Barique MA, Ohigashi H. Polymer 2001;42(11):4981e7.
[72] Guan F, Wang J, Yang L, Tseng JK, Han K, Wang Q, et al. Macromolecules
2011;44(7):2190e9.
[73] Guan F, Wang J, Yang L, Guan B, Han K, Wang Q, et al. Advanced Functional
Materials 2011;21(16):3176e88.
[74] Kepler RG. Ferroelectric, pyroelectric, and piezoelectric properties of
poly(vinylidene uoride). In: Nalwa HS, editor. Ferroelectric polymers:
chemistry, physics, and applications. New York: Marcel Dekker, Inc.; 1995.
p. 183e232 [Chapter 3].
[75] Bur AJ, Barnes JD, Wahlstrand KJ. Journal of Applied Physics 1986;59(7):
2345e54.
[76] Lovinger AJ, Cais RE. Macromolecules 1984;17(10):1939e45.
[77] Su R, Tseng JK, Lu MS, Lin M, Fu Q, Zhu L. Polymer 2012;53(3):728e39.
[78] Li GR, Kagami N, Ohigashi H. Journal of Applied Physics 1992;72(3):1056e61.
[79] Kremer F, Schnhals A. Broadband dielectric spectroscopy. New York:
Springer; 2003.
[80] Mackey M, Schuele DE, Zhu L, Baer E. Journal of Applied Physics 2012;
111(11):113702.
[81] Wang JL, Yuan SZ, Tian L, Meng XJ, Sun JL, Chu JH. Applied Physics Letters
2011;98(5):052906.
[82] Kobayashi M, Tashiro K, Tadokoro H. Macromolecules 1975;8(2):158e71.
[83] Tashiro K, Kobayashi M, Tadokoro H. Macromolecules 1981;14(6):1757e64.
[84] Bachmann MA, Gordon WL, Koenig JL, Lando JB. Journal of Applied Physics
1979;50(10):6106e12.
[85] Mackey M, Schuele DE, Zhu L, Flandin L, Wolak MA, Shirk JS, et al.
Macromolecules 2012;45(4):1954e62.
[86] Shilov SV, Skupin H, Kremer F, Gebhard E, Zentel R. Liquid Crystals 1997;
22(2):203e10.
[87] Lu Y, Claude J, Neese B, Zhang Q, Wang Q. Journal of the American Chemical
Society 2006;128(25):8120e1.
[88] Xu H, Shen D, Zhang Q. Polymer 2007;48(7):2124e9.
[89] Lovinger AJ. Macromolecules 1985;18(5):910e8.
[90] Lovinger AJ. Radiation effects on the structure and properties of poly(vinylidene
uoride) and its ferroelectric copolymers. In: Clough RL, Shalaby W, editors.
Radiation effects on polymers. Washington, DC: American Chemical Society;
1991. p. 84e100.
[91] Tang Y, Wang D, Guo S, Zhao XZ. European Polymer Journal 2001;37(3):
471e4.
[92] Welter C, Faria LO, Moreira RL. Physical Review B 2003;67(14):144103.
[93] Lam TY, Lau ST, Chan HLW, Choy CL, Cheung WY, Wong SP. Journal of
Polymer Science, Part B: Polymer Physics 2005;43(17):2334e9.
[94] Lau ST, Chan HLW, Choy CL. Applied Physics A: Materials Science &
Processing 2005;80(2):289e94.
[95] Lau ST, Leung KY, Chan HLW, Choy CL, Sundaravel B, Wilson I. Ferroelectrics
2002;273(1):2387e92.
[96] Mabboux PY, Gleason KK. Journal of Fluorine Chemistry 2002;113(1):27e35.
[97] Wang L, Zhao X, Feng J. Journal of Polymer Science, Part B: Polymer Physics
2006;44(12):1714e24.
[98] Cheng ZY, Olson D, Xu H, Xia F, Hundal JS, Zhang QM, et al. Macromolecules
2002;35(3):664e72.
[99] Sumpter BG, Noid DW, Liang GL, Wunderlich B. Atomistic Modeling of
Physical Properties 1994;116:27e72.
[100] Li ZM, Arbatti MD, Cheng ZY. Macromolecules 2004;37(1):79e85.
[101] Li ZM, Li SQ, Cheng ZY. Journal of Applied Physics 2005;97(1):014102.
[102] Byutner OG, Smith GD. Macromolecules 2000;33(11):4264e70.
[103] Mazars M. Journal of Chemical Physics 2001;115(7):2955e65.
[104] Scherowsky G. Ferroelectric liquid crystal (FLC) polymers. In: Nalwa HS,
editor. Ferroelectric polymers: chemistry, physics, and applications. New
York: Marcel Dekker, Inc.; 1995. p. 35e538 [Chapter 10].
[105] Blythe AR, Bloor D. Electrical properties of polymers. 2nd ed. Cambridge;
New York: Cambridge University Press; 2005.
[106] Black CT, Farrell C, Licata TJ. Applied Physics Letters 1997;71(14):2041e3.
[107] Riande E, Daz-Calleja R. Electrical properties of polymers. New York: Marcel
Dekker; 2004.
[108] Guan F, Wang J, Pan J, Wang Q, Zhu L. Chinese Journal of Polymer Science
2011;29(1):65e80.
[109] Furukawa T. Advances in Colloid and Interface Science 1997;71-72:183e208.
[110] Furukawa T, Date M, Fukada E. Journal of Applied Physics 1980;51:1135e41.
[111] Yamada T, Ueda T, Kitayama T. Journal of Applied Physics 1981;52(2):948e52.
[112] Furukawa T, Lovinger AJ, Davis GT, Broadhurst MG. Macromolecules 1983;
16(12):1885e90.
[113] Koizumi N, Murata Y, Tsunashima H. IEEE Transactions on Electrical
Insulation 1986;21(3):543e8.
[114] Takahashi Y, Kodama H, Nakamura M, Furukawa T, Date M. Polymer Journal
1999;31(3):263e7.
[115] Furukawa T, Takahashi Y. Ferroelectrics 2001;264(1e4):1739e48.
[116] Guan F, Yuan Z, Shu EW, Zhu L. Applied Physics Letters 2009;94(5):052907.
[117] Zhang M, Russell TP. Macromolecules 2006;39(10):3531e9.
[118] Enkelmann V. Advances in Polymer Science 1984;63:91e136.
[119] Lauher JW, Fowler FW, Goroff NS. Accounts of Chemical Research 2008;
41(9):1215e29.

1728

L. Yang et al. / Polymer 54 (2013) 1709e1728


Lianyun Yang was born in Jilin Province, China. He
obtained his B.S. degree in polymer chemistry at University
of Science and Technology of China in 2009. Currently, he
is a Ph.D. candidate in the Department of Macromolecular
Science and Engineering at Case Western Reserve University. His current research interest focuses on novel ferroelectric polymers as high-k/low-loss dielectrics for energy
storage applications.

Xinyu Li obtained a B.E. degree in Department of Materials


Science and Engineering, Tsinghua University, Beijing,
China in 2009. Presently, he is pursuing a Ph.D. degree in
the Department of Electrical Engineering at the Pennsylvania State University. His current research focuses on the
development of organic materials exhibiting novel electrical properties, especially giant electrocaloric response,
as well as the design of solid-state cooling devices based
on the electrocaloric effect.

Elshad Allahyarov is currently a researcher at the Physics


Department of Case Western Reserve University. He also
holds scientic positions as a researcher at the Institute for
Theoretical Physics II in Heinrich-Heine University of Dusseldorf, Germany, and as a senior researcher in the Theoretical Department of the Institute for High Temperatures
of Russian Academy of Sciences in Moscow, Russia. He
received his M.S. degree in Theoretical Physics in 1988
from Moscow State University, Russia. He obtained his
Ph.D. degree in Theoretical Physics in 1995, and his Doctor
of Sciences (Habilitation) degree in Theoretical Physics in
2005, both from Russian Academy of Sciences. His research
interest includes the physics of soft matter, nucleation
problems in dense systems, morphological changes in ionomers under external elds,
the role of Coulomb correlations in colloidal and biophysical systems, and energy storage in charged polymers.

Philip L. Taylor is Distinguished University Professor and


Perkins Professor of Physics and Macromolecular Science
and Engineering at Case Western Reserve University. He
received his B.Sc. degree in Physics in 1959 from Kings
College, London, and his Ph.D. in Theoretical Physics in
1963 from the University of Cambridge. He has held visiting positions at the Universities of Bonn, Manchester,
Utrecht, Cambridge, Oregon, and Washington. He is a Fellow of the American Physical Society and of the American
Association for the Advancement of Science. His research
interests include theoretical and simulation studies of the
electrical and thermodynamic properties of polymers and
liquid crystals.

Qiming Zhang is currently Distinguished Professor of


Electrical Engineering and Materials Science and Engineering at Penn State University. The research areas in his
group include fundamentals and applications of novel
electronic and electroactive materials. Research activities
in his group cover actuators, sensors, transducers, dielectrics and charge storage devices, polymer thin lm
devices, polymer MEMS, electrocaloric effect and solidstate cooling devices and electro-optic and photonic devices. He has over 340 publications and 10 patents in these
areas. He is the recipient of the 2008 Penn State Engineering Society Premier Research Award and a Fellow of IEEE
and a Fellow of APS.

Lei Zhu is currently an Associate Professor of Macromolecular Science and Engineering at Case Western
Reserve University. He received his B.S. degree in Materials
Chemistry in 1993 and M.S. degree in Polymer Chemistry
and Physics in 1996 from Fudan University. He earned his
Ph.D. degree in Polymer Science in 2000 from the University of Akron. After two-year post-doctoral experience
at University of Akron, he joined Institute of Materials Science and Department of Chemical, Materials and Biomolecular Engineering at University of Connecticut, as an
Assistant Professor. In 2009, he moved to Case Western
Reserve University. He is recipient of National Science
Foundation (NSF) Career Award, 3M Non-tenured Faculty
Award, and DuPont Young Professor Award. His research interests include high-k polymer and polymereinorganic hybrid materials for electrical applications, development
of articial antibody as nanomedicines, and supramolecular self-assembly of supermolecules and polymers. He is author and co-author of w100 refereed journal publications and 5 book chapters.

You might also like