Вы находитесь на странице: 1из 8

This article appeared in a journal published by Elsevier.

The attached
copy is furnished to the author for internal non-commercial research
and education use, including for instruction at the authors institution
and sharing with colleagues.
Other uses, including reproduction and distribution, or selling or
licensing copies, or posting to personal, institutional or third party
websites are prohibited.
In most cases authors are permitted to post their version of the
article (e.g. in Word or Tex form) to their personal website or
institutional repository. Authors requiring further information
regarding Elsevier’s archiving and manuscript policies are
encouraged to visit:
http://www.elsevier.com/copyright
Author's personal copy

European Polymer Journal 48 (2012) 79–85

Contents lists available at SciVerse ScienceDirect

European Polymer Journal


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

Dielectric spectroscopy of biodegradable


poly(3-hydroxybutyrate-co-3-hydroxyhexanoate) films
Yi Yang a,b, Shanming Ke b, Li Ren a,⇑, Yingjun Wang a, Yiyang Li c, Haitao Huang a,b,⇑
a
National Engineering Research Center for Tissue Regeneration and Reconstruction, School of Materials Science and Engineering,
South China University of Technology, Guangzhou 510641, China
b
Department of Applied Physics and Materials Research Center, The Hong Kong Polytechnic University, Hung Hom, Kowloon, Hong Kong
c
Department of Mechanical and Automation Engineering, The Chinese University of Hong Kong, Sha Tin, New Territories, Hong Kong

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

Article history: In this paper, we report a systematic study of the dielectric relaxation spectroscopy of bio-
Received 21 June 2011 degradable poly(3-hydroxybutyrate-co-3-hydroxyhexanoate) (PHBHHx) polyester which
Received in revised form 15 August 2011 has potential applications as a ‘‘green’’ dielectric material in electronic devices. The dielec-
Accepted 7 October 2011
tric spectra was measured over a wide frequency range (100  107 Hz) from 100 to 60 °C.
Available online 15 October 2011
A glass and a sub-glass transition relaxations were observed in the dielectric spectra of
PHBHHx. In addition, a nearly constant loss behavior was found by analyzing the dielectric
Keywords:
and conductivity spectra.
Dielectric spectroscopy
PHBHHx
Ó 2011 Elsevier Ltd. All rights reserved.
Biodegradable polyester
Green dielectric materials
Nearly constant loss

1. Introduction tensile strength of 20 MPa, while PHB film only has an elon-
gation at break of less than 10% [7,8]. Previous studies also
Polyhydroxyalkanoates (PHA) are a bio-polyester family showed that PHBHHx has better biocompatibilities with
which is produced by bacteria under unbalanced growth [1]. fibroblast, chondrocyte and nerve cells than PHB [5,9].
This family of materials has attracted more and more atten- The glass transition temperature (Tg  0 °C for PHBHHx
tion in tissue engineering and other biomedical applications with 12 mol% HHx) and melting temperature (Tm) of
due to their microbial synthesis [2], good biocompatibility, PHBHHx vary with the number of the HHx units, which have
biodegradability and promoting cell growth [3–5]. Poly(3- been well investigated by differential scanning calorimeter
hydroxybutyrate-co-3-hydroxyhexanoate) (PHBHHx) is a (DSC) and thermogravimetric analysis (TGA) [10,11]. Re-
kind of copolyesters consisting of 3-hydroxybutyrate (HB) cently, PHBHHx has been produced in a large scale [12],
and randomly distributed 3-hydroxyhexanoate (HHx) units which is very important for practical applications. Apart
[6]. As a common member of the PHA family, PHBHHx from a variety of biological and medical applications,
shows better thermal and mechanical properties, such as PHBHHx may find its potential applications as a ‘‘green’’
flexibility and thermoplasticity, than those of many other dielectric material for capacitors [13]. Capacitors with high
PHAs. For instance, PHBHHx (with 12 mol% HHx) has an energy storage density are becoming increasingly important
elongation at break of 500% and maintains an applicable to a broad range of applications, especially in pulsed power
circuit applications, such as implantable medical devices.
The term ‘‘green’’ in electronics actually refers to lead- and
⇑ Corresponding authors at: Department of Applied Physics and Mate- halogen-free materials. To meet the ever rising require-
rials Research Center, The Hong Kong Polytechnic University, Hung Hom,
ments on environment protection, a new generation of
Kowloon, Hong Kong (H. Huang).
E-mail addresses: psliren@scut.edu.cn (L. Ren), aphhuang@polyu.
‘‘green’’ dielectrics has to be composed of biodegradable
edu.hk (H. Huang). polymers which can be decomposed when buried, without

0014-3057/$ - see front matter Ó 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.eurpolymj.2011.10.002
Author's personal copy

80 Y. Yang et al. / European Polymer Journal 48 (2012) 79–85

affecting the environment and ecosystem. PHBHHx is 2.5. Dynamic mechanical analysis (DMA)
biodegradable and it is worthwhile to study its dielectric
properties for potential applications as ‘‘green’’ dielectric The glass transition temperature (Tg) of the pure PHBHHx
material. Our previous study [13] on PHBV has shown that polymer was determined by dynamic mechanical analysis
PHBV has a higher dielectric constant (e0 = 8.5 for 10 KHz (DMA). A Perkin Elmer Diamond DMA was used in tensile
at room temperature) than that of the conventional dielec- mode at a frequency of 1 Hz. The sample was heated from
tric polymer, such as polypropylene (PP, e0 = 2 for 10 KHz 100 to 100 °C at a rate of 5 °C/min.
at room temperature) and polycarbonate (PC, e0 = 2.5 for
10 KHz at room temperature). The electrical energy density 3. Results and discussion
stored in PHBV is also much higher than that stored in PP
and PC. This leads us to give the first detailed study on the 3.1. 1H NMR and SEM
dielectric properties of PHBHHx by using the dielectric
relaxation spectroscopy (DRS) which is one of the most The chemical structure of PHBHHx, as shown in the in-
suitable and versatile techniques to assess the dynamics of set of Fig. 1, contains [–O–CH(CH3)–CH2–CO–] for HB and
polymer materials in various time and length scales. [–O–CH(C3H7)–CH2–CO–] for HHx. As characterized by
the 1H NMR spectrum in Fig. 1, a strong proton signal at
2. Experimental 1.26 ppm can be attributed to –CH3 in HB while another
methyl group in HHx appeared at 0.91 ppm. The signal
2.1. Materials and film preparation appearing at 1.57 ppm can be assigned to the H atom res-
onances of the –CH2–CH2– unit in HHx. All these results
Poly(3-hydroxybutyrate-co-3-hydroxyhexanoate) (PHB are consistent with previous reports [2,14], indicating that
HHx, with HHx content of 12%) raw powders were used as pure PHBHHx was obtained.
received from Jiangmen Center for Biotechnology Develop-
ment, Guangdong, China. The PHBHHx films were fabricated 3.2. DMA
by hot-pressing. The raw powders were compression-
molded directly under a pressure of 20 MPa at 120 °C for Similar to PHB, PHBHHx is a semi-crystalline polymer
2 h. The PHBHHx films with a thickness of 0.05 mm were with a melting point Tm  124 °C [15]. From the X-ray
then obtained and cut into 10  10 mm2 pieces for dielectric diffraction patterns (not shown here), the crystallinity (Xc)
measurements. of the PHBHHx films was estimated to be 7.6%. The low de-
gree of crystallinity will leads to some interesting relaxation
2.2. Nuclear magnetic resonance spectroscopy (NMR) processes which different from the crystalline polymers
[16,17]. DMA was also performed to investigate the
1H Nuclear magnetic resonance (1H NMR) spectroscopy thermomechanical properties of PHBHHx films. The storage
of PHBHHx was recorded on a Bruker AG 300 NMR Spec- modulus (G0 ), loss modulus (G00 ) and tan d plots over a wide
trometer in CDCl3 at room temperature. Tetramethylsilane temperature range are presented in Fig. 2. At temperatures
was used as the internal reference. higher than 0 °C, the storage modulus decreases, manifest-
ing a rubbery state of the PHBHHx polyester. The accompa-
nied relaxation is associated to the apparent glass transition
2.3. Scanning electron microscopy (SEM)
and Tg can be defined as the temperature corresponding to
The morphology of the obtained films was examined
under a scanning electron microscope (SEM, JEOL JSM-
6490). Before observation, the specimen was fractured in
liquid nitrogen and coated with gold. The acceleration volt-
age of the SEM apparatus was controlled at 20 kV. Thick-
ness of the PHBHHx film was measured from the SEM
image and also by a vernier caliper.

2.4. Dielectric relaxation spectroscopy (DRS)

For dielectric measurements, silver was sputtered on


both surfaces of the specimens. The dielectric properties
and ac conductivity were measured by using a broadband
frequency-response analyzer (Alpha analyzer, Novocontrol
Technologies) over a frequency range of 100–107 Hz at var-
ious temperatures. The temperature was controlled by a
Novocontrol Quatro Cryosystem with high accuracy
(±0.1 K). The real and imaginary parts of dielectric permit-
tivity [e⁄(f) = e0 (f)  ie00 (f), where f is the probe frequency]
were collected isothermally in the order of increasing Fig. 1. 300 MHz 1H NMR spectrum of PHBHHx in CDCl3. The inset shows
temperatures. the chemical structure of PHBHHx.
Author's personal copy

Y. Yang et al. / European Polymer Journal 48 (2012) 79–85 81

Fig. 2. Dynamical mechanical analysis results: storage (G0 ) and loss (G00 )
modulus, and tan d of the PHBHHx film.
Fig. 3. Temperature dependence of the (a) real and (b) imaginary parts of
the relative dielectric permittivity of PHBHHx at selected frequencies.
the maximum of either the loss modulus or the tan d. The
values of Tg are 8.9 and 12 °C, respectively, as determined
from the loss modulus and the tan d. A weak cold crystalliza-
tion peak could be observed at Tcc  87.8 °C. These are con- glass transition is still an open question and it has been
assigned to the reorientation of the ester groups [22] or re-
sistent with previous reports [15].
lated to the torsional vibrations of main chains in the
neighborhood of its local equilibrium conformation [23].
3.3. Dielectric relaxation spectroscopy It should be noted that the b process in PHBHHx merges
into an extremely broad peak with increasing probe fre-
As a new member of the PHA family, PHBHHx still has quency and there is no well-defined peak maximum above
many properties that are interesting for a wide range of 1 KHz, as shown in Fig. 3b. This result implies that there
applications. In this section, we provide detailed dielectric should be other contributions to PHBHHx in this tempera-
studies to extend the research area of biodegradable poly- tures range, which dominates the high frequency DRS.
esters and to give a deeper understanding on the dielectric At temperatures above 50 °C (region II), another relax-
properties of PHBHHx. ation process becomes evident. This process is very similar
Depending on the structure, polymer molecules may to a relaxation which is related to the glass transition in
show different kinds of molecular motions leading to dif- polymers and the segmental mobility. The peak height of
ferent relaxations. The temperature dependences of the the a relaxation in PHBHHx seems unchanged with
real (e0 ) and the imaginary (e00 ) parts of the relative dielec- increasing frequencies. It is interesting that this relaxation
tric permittivity of PHBHHx films at selected frequencies process took place over a wide temperature range, cover-
are shown in Fig. 3. Similar to other biodegradable polyes- ing the glass transition temperature obtained from the
ter (such as Poly(butylene succinate), PBS) [18,19], three DMA results (Tg  8.9–12 °C) [16]. The a relaxation peak
types of dielectric relaxation processes can be observed: in PHBHHx became broader and the peak height was al-
region I (at low temperatures, 100 to 50 °C), region II most unchanged with increasing frequencies. This is quite
(at intermediate temperatures, 50 to 60 °C), and region different from PBS, whose peak height shows an intense
III (at high temperatures, 20–60 °C). In region I, a broad rising trend with increasing frequencies. This implies that
relaxation peak in e00 (T) can be seen to move toward higher the segmental mobility in PHBHHx is much lower than
frequencies as the temperature increases, which is accom- that in PBS which reflects its viscous nature. The space
panied by a slow transition step in the e0 curve. This has charge effect [24] is evident in region III, where e0 and e00 in-
been identified as a sub-glass b transition process similar crease dramatically at low frequencies with increasing
to that in many other hyperbranched polymersexplored temperatures (Fig. 3).
by using dielectric relaxation spectroscopy [19,20] and dy- The frequency dependences of the relative dielectric
namic mechanical analysis [21]. The origin of such sub- permittivity (both e0 and e00 ) for a and b relaxations are
Author's personal copy

82 Y. Yang et al. / European Polymer Journal 48 (2012) 79–85

Fig. 4. Frequency dependence of the (a) real and (b) imaginary parts of Fig. 5. Frequency dependence of the (a) real and (b) imaginary parts of
the relative dielectric permittivity of PHBHHx at around the a glass the relative dielectric permittivity of PHBHHx at around the b sub-glass
transition temperature. The dots are experimental data and the solid transition temperature. The dots are experimental data and the solid
curves are the best-fitted curves according to Eqs. (2) and (3). curves are the best-fitted curves according to Eqs. (2) and (3).

plotted in Figs. 4 and 5, respectively. For quantitative eval- dielectric loss peak corresponding to the b relaxation, as
uation of the relaxation time s(T) and other relaxation shown in Fig. 3(b).
parameters, the data were fitted by a Cole–Cole dielectric
equation [25,26],
3.4. Mapping of the relaxation parameters
 De
e ðxÞ ¼ e1 þ ; 0 < a P 1; ð1Þ
1 þ ðixsc Þ1a The relaxation shape parameter a and the characteristic
time sc were evaluated from the frequency spectra of the
  dielectric permittivity by using the Cole–Cole functions
De sinhðbuÞ
e0 ðxÞ ¼ e1 þ 1 ð2Þ (Eqs. (1)–(3)). The evolution of a and the distribution of
2 coshðbuÞ þ cosðbp=2Þ
sc (for both the a and b relaxations) are shown in Figs. 6
and 7 as a function of temperature. Clearly, the shape
De sinðbp=2Þ parameter a associated with the a relaxation increases
e00 ðxÞ ¼ ; b ¼ 1  a; ð3Þ
2 coshðbuÞ þ cosðbp=2Þ from 0.4 to 0.7 with decreasing temperatures, while it
maintains at a fixed value of around 0.7 for the b relaxa-
where u = ln(xsc), x is the angular frequency, sc is the
tion. Larger a means a much broader distribution of relax-
characteristic time, a is the shape parameter, De = es  e1
ation times [25]. For a relaxations, the distribution of
is the dielectric relaxation strength with es and e1 defined
relaxation time becomes broader upon cooling, which
as the low and high frequency limits of the relative dielec-
can be better viewed from Fig. 6b. Here, the distribution
tric permittivity. It can be seen from Figs. 4 and 5 that the
of relaxation time derived from the Cole–Cole relation
low frequency permittivity (i.e. <100 Hz) at 0 °C does not
can be expressed as [25],
follow the fitting curves, which was probably due to the
space charge polarization [24]. For the b relaxation, which sinðapÞ
is visible at relatively low frequency (Fig. 3), the corre- f ðsÞ ¼ ð4Þ
2p½coshðð1  aÞ lnðsc =sÞÞ  cosðapÞ
sponding e00 (f) becomes almost frequency independent at
higher frequencies (as shown in Fig. 5b), similar to a where f(s) can be calculated from the fitting parameters of
‘‘nearly constant loss’’ (NCL) process [27]. This ‘‘NCL’’ the dielectric permittivity.
behavior will be discussed in Section 3.5 and is obviously It can also be seen from Fig. 6b that the distribution of
the main reason for the flattening of the high frequency relaxation times for b relaxation is nearly temperature
Author's personal copy

Y. Yang et al. / European Polymer Journal 48 (2012) 79–85 83

[22]. Due to the large free spaces in polymers, the freezing


of ester groups may not influence the distribution of relax-
ation times.
The temperature dependences of the characteristic time
sc for the a and b relaxations are shown in Fig. 7. Generally,
the relaxation for a sub-glass transition process always fol-
lows an Arrhenius law. The characteristic time for the
relaxation process in region II of Fig. 3 does not follow an
Arrhenius law exactly (Fig. 7a), but can be well described
by the Vögel–Fulcher (VF) equation [26,28]:
1 Ea
sc ¼ exp ð5Þ
x0 kB ðT  T f Þ
where Ea is the activation energy, x0 is the attempt fre-
quency of a molecule and Tf is the static freezing tempera-
ture at which x goes to zero. This behavior indicates that
the relaxation process in region II is a genuine a process.
The large difference between the viscoelastic and dielectric
peak may be understood by its semi-crystalline structure
with very low degree of crystallinity [16,17]. It should be
emphasized that Ea obtained from VF relation is a parame-
ter corresponding to the hindering barrier of the Arrhenius
case (i.e. the case for Tf = 0 K), but it does not represent a
real activation process. The calculated activation energy
Ea for the a relaxation is 91.9 meV, while the freezing tem-
perature Tf is 212 K. Besides, the relationship between the
probe frequency f and the peak temperature Tm obtained
from Fig. 3b for the glass transition a is also shown in
Fig. 7 and can be described by the following VF equation,
Ea
f ¼ f0 exp ð6Þ
kB ðT  T f Þ
The obtained activation energy and freezing tempera-
Fig. 6. (a) Temperature dependence of a obtained from the fitting of data
ture are 90.2 meV and 211 K, respectively, and are in good
shown in Figs. 4 and 5. (b) Cole–Cole distribution of relaxation times at
selected temperatures. agreement with the results obtained from the sc–T relation.
Fig. 7c shows the Arrhenius plot for the b relaxation. The
activation energy for this process is 679 meV.
independent. This feature is helpful to understanding the
origin of the b sub-glass transition process. However, it 3.5. Nearly constant loss-like behavior
cannot lead to the conclusion that the b sub-glass transi-
tion process is related to the torsional vibrations of main At low temperatures, there are two obvious contribu-
chains in the neighborhood of its local equilibrium confor- tions to DRS of PHBHHx (Fig. 6): one is the sub-glass tran-
mation [23], or to the reorientation of the ester groups sition at low frequency and the other is the ‘‘nearly

Fig. 7. Temperature dependence of the characteristic relaxation time sc (obtained from the fitting of data shown in Figs. 4 and 5) for the a and b transitions.
(a) and (c) represent the Arrhenius plots. Also shown in (b) is the relation between the probe frequency and the temperature corresponding to the
maximum dielectric permittivity (imaginary part) of the a relaxation in PHBHHx. The short dashed lines in (a) just guide the eyes. The solid and dashed
lines in (b) are fitted by the Vögel–Fulcher law. The solid line in (c) is fitted by the Arrhenius law.
Author's personal copy

84 Y. Yang et al. / European Polymer Journal 48 (2012) 79–85

constant loss’’-like (NCL) process. Actually, NCL is a univer- temperature. (It should be noted that the NCL-like response
sal behavior in disordered materials (such as glasses, poly- above 5 °C is out of our experimental frequency window.)
mers and highly doped crystals) [29–31]. At sufficiently This behavior is very similar to that of ionic conducting
low temperatures and/or high frequencies, the conductiv- glasses, such as CeO2:Gd2O3 [27], but is quite different from
ity r0 (f) for NCL response depends linearly on frequency, that of ferroelectric ceramics [32] and some other biode-
which implying a negligible frequency dependence of the gradable polyester, such as poly(butylene succinate) [18],
imaginary part of dielectric permittivity [e00 (f) = r0 (f)/2pf]. whose s increases linearly with increasing temperature
The frequency dependence of the conductivity is given by leading to a superlinear power-law behavior [33].
From the Kramers–Krönig relation [26,32,34], it is
r0 ðf Þ ¼ Af s ð7Þ known that the power law (2pf)n of the ac conductivity will
lead to a (2pf)n1 behavior for the real part of the dielectric
where the exponent s is nearly equal to 1 and A is a con- permittivity, e.g.,
stant with weak temperature dependence [29].
Fig. 8a–c show the frequency dependence of the con-
e0 ðf Þ ¼ Bð2pf Þn1 ð8Þ
ductivity r0 (f) and imaginary part of permittivity e00 (f) at se-
lected temperatures. At low temperature (95 °C), r0 (f) 0
We plotted the frequency dependence of e (f) at selected
exhibits a linear frequency dependence above 200 Hz (with temperatures in Fig. 9a. Clearly, the e0 (f) at the NCL-like re-
an exponent s = 1 following Eq. (7)) and therefore a negli- gime could be described well by Eq. (8), which again
gible frequency dependence of e00 , an NCL-like behavior. reflecting a genuine NCL behavior. The temperature depen-
At 45 °C, the right wing of b process is absolutely covered dence of the exponent n displays a very similar behavior to
by the NCL-like response and meanwhile the a process sets that of exponent s, as shown in Fig. 9(b).
in at low frequencies. The span of the NCL-like regime Although the apparent dielectric response behaves as
becomes narrower with increasing temperatures. constant loss, it should be emphasized that we cannot con-
By fitting Eq. (7), the exponent s can be obtained and its clude here an unambiguous NCL behavior in PHBHHx.
temperature dependence is shown in Fig. 8d. Two regions Overlapping of two or several relaxation processes could
can be clearly identified. Below 25 °C, the exponent s also give rise to a ‘‘nearly constant loss’’. Further experi-
reaches a plateau value of 1, while at a higher temperature mental investigation is needed to confirm our finding is a
(25  5 °C), s decreases quickly with increasing real NCL response or multi relaxation overlapping.

Fig. 8. (a, b, c) Frequency dependence of the conductivity and the imaginary part of the dielectric permittivity at selected temperatures for PHBHHx. The
solid lines are the best fits according to Eq. (7). (d) Temperature dependence of the exponent s from the fitting data.
Author's personal copy

Y. Yang et al. / European Polymer Journal 48 (2012) 79–85 85

Acknowledgements

This work was supported by National Natural Science


Foundation of China (Grant Nos. 50732003, 50803018
and U0834003), National Science and Technology Support
Programme of China (Grant No. 2006BAI16B06), the Hong
Kong Polytechnic University (Project Nos. A-PK30 and A-
PJ46) and Endowment Fund Research Grant of United Col-
lege, The Chinese University of Hong Kong (Project No.
CA11168).

References

[1] Gerco P, Martuscelli E. Polymer 1989;30:1475–83.


[2] Li ZB, Yang XD, Wu LP, Chen ZF, Lin YT, Xu KT, et al. J Biomater Sci
2009;20:1179–202.
[3] Chen GQ, Wu Q. Biomaterials 2005;26:6565–78.
[4] Lee SY. Biotechnol Bioeng 1996;49:1–14.
[5] Yang X, Zhao K, Chen GQ. Biomaterials 2002;23:1391–7.
[6] Xu J, Guo BH, Yang R, Wu Q, Chen GQ, Zhang ZM. Polymer
2002;43:6893–9.
[7] Zhao K, Deng Y, Chen JC, Chen GQ. Biomaterials 2003;24:1041–5.
[8] Alata H, Aoyama T, Inoue T. Macromolecules 2007;40:4546–51.
[9] Deng Y, Zhao K, Zhang XF, Hu P, Chen GQ. Biomaterials
2002;23:4049–56.
[10] Doi Y, Kitamura S, Abe H. Macromolecules 1995;28:4822–8.
[11] Asrar J, Valentin HE, Berger PA, Tran M, Padgette SR, Garbow JR.
Biomacromolecules 2002;3:1006–12.
[12] Chen GQ, Zhang G, Park SJ, Lee SY. Appl Microbiol Biotechnol
2001;57:50–5.
[13] Ke SM, Huang HT, Ren L, Wang YJ. J Appl Phys 2009;105:096103.
[14] Pan JY, Li GY, Chen ZF, Chen XY, Zhu WF, Xu KT. Biomaterials
2009;30:2975–84.
[15] Ye HM, Wang Zhen, Wang HH, Chen GQ, Xu J. Polymer
2010;51:6037–46.
Fig. 9. (a) Frequency dependence of the real part of the relative dielectric
[16] Kremer F, Schonhals A. Broadband dielectric
permittivity at selected temperatures. Dots are experimental values and
spectroscopy. Berlin: Springer; 2003.
the solid lines are the best fits according to Eq. (8). (b) Temperature [17] Tai HJ. Polymer 2007;48:4558–66.
dependence of the exponent n and constant B fitted according to Eq. (8). [18] Yu L, Ke SM, Zhang YH, Huang HT. Unpublished work.
[19] Sen S, Boyd RH. Eur Polym J 2008;44:3280–7.
[20] Bravard SP, Boyd RH. Macromolecules 2003;36:741–8.
4. Conclusions [21] Ray SS, Okamoto K, Okamoto M. Macromolecules 2003;36:2355–67.
[22] Malmstrom E, Hult A, Gedde UW, Liu F, Boyd RH. Polymer
1997;38:4873–9.
In summary, we have performed a systematic dielectric
[23] Tatsumi T, Ito E, Hayakawa R. J Polym Sci B Polym Phys
study on the biodegradable polyester PHBHHx. The sub- 1992;30:701–6.
glass transition relaxation (a and b) and nearly constant [24] Tai HJ. Polymer 2008;49:2328–33.
[25] Cole KS, Cole RH. J Chem Phys 1941;9:341–51.
dielectric loss behavior in PHBHHx have been addressed
[26] Ke SM, Fan HQ, Huang HT. Appl Phys Lett 2010;97:132905.
and discussed. The a process shows a Vögel–Fulcher tem- [27] Lee WK, Liu JF, Nowick AS. Phys Rev Lett 1991;67:1559–61.
perature dependence, while the b relaxation follows the [28] Viehland D, Jang SJ, Cross LE, Wuttig M. J Appl Phys
Arrhenius law. The distribution of relaxation times is sym- 1990;68:2916–21.
[29] Rivera A, León C, Varsamis CPE, Chryssikos GD, Ngai KL, Roland CM,
metric for both a and b processes. However, the width of et al. Phys Rev Lett 2002;88:125902.
distribution is temperature independent for b process, [30] Sidebottom DL, Murray-Krezan CM. Phys Rev Lett 2002;89:195901.
but becomes broader upon cooling for a process. The cou- [31] Roling B, Martiny C, Murugavel S. Phys Rev Lett 2001;87:085901.
[32] Ke SM, Huang HT, Yu SH, Zhou LM. J Appl Phys 2010;107:084112.
pling between the b relaxation and NCL-like response in [33] Lunkenheimer P, Loidl A. Phys Rev Lett 2003;91:207601.
PHBHHx is interesting and is may be useful for a plausible [34] Jonscher AK. Dielectric relaxation in solids. London: Chelsea
explanation of NCL. Dielectric; 1983.

Вам также может понравиться