You are on page 1of 17

Journal of Macromolecular Science, Part B

Physics

ISSN: 0022-2348 (Print) 1525-609X (Online) Journal homepage: http://www.tandfonline.com/loi/lmsb20

Effects of Molecular Weight on Thermal


Degradation of Poly(-methyl styrene) in Nitrogen

Lina Ye, Meifang Liu, Yawen Huang, Zhanwen Zhang & Junxiao Yang

To cite this article: Lina Ye, Meifang Liu, Yawen Huang, Zhanwen Zhang & Junxiao Yang (2015)
Effects of Molecular Weight on Thermal Degradation of Poly(-methyl styrene) in Nitrogen, Journal
of Macromolecular Science, Part B, 54:12, 1479-1494, DOI: 10.1080/00222348.2015.1094645

To link to this article: http://dx.doi.org/10.1080/00222348.2015.1094645

Accepted author version posted online: 05


Oct 2015.
Published online: 05 Oct 2015.

Submit your article to this journal

Article views: 60

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=lmsb20

Download by: [Indian Institute of Technology Madras] Date: 19 July 2017, At: 22:33
Journal of Macromolecular Science , Part B: Physics, 54: 14791494, 2015
Copyright Taylor and Francis Group, LLC
ISSN: 0022-2348 print / 1525-609X online
DOI: 10.1080/00222348.2015.1094645

Effects of Molecular Weight on Thermal


Degradation of Poly(a-methyl styrene) in Nitrogen

LINA YE,1,2 MEIFANG LIU,1 YAWEN HUANG,2 ZHANWEN


ZHANG,1 AND JUNXIAO YANG2
1
Research Center of Laser Fusion, China Academy of Engineering Physics,
Mianyang, China
2
School of Material Science and Engineering, Southwest University of Science
and Technology, Mianyang, China

The effects of molecular weight on the thermal degradation behavior of poly(a-methyl


styrene) (PAMS) was investigated by pyrolysis-gas chromatography-mass
spectrometry (Py-GC/MS) and thermogravimetric analysis (TGA). The Py-GC/MS
analysis results showed that the degradation of PAMS with different molecular
weights in nitrogen produced only the monomer, alpha-methylstyrene. The TGA
results showed a pronounced reduction in the decomposition temperature with
increasing molecular weight. The degradation kinetic parameters, calculated by the
Kissinger and the CoatsRedfern methods, further revealed that the activation energy
and the pre-exponential factor decreased with increasing molecular weight. Most
importantly, the degradation order of the PAMS in nitrogen remained around 1,
independent of the molecular weight, suggesting the maintenance of the
depolymerization mechanism. All the above results provided an insight into the effects
of molecular weight on the thermal degradation behavior of PAMS.

Keywords degradation kinetic, inertial confinement fusion, molecular weight, poly


(a-methyl styrene)

Introduction
Polymer degradation is an important practical topic in polymer science and engineering
that pertains to their applications, longevity, recycling and so on.[1,2] Taking advantage of
polymer degradation, the decomposable mandrels approach to form plastic mandrels
has been developed for preparing glow plasma polymer (GDP) shells used in the inertial
confinement fusion (ICF) experiments.[38] A spherical mandrel is first prepared from
poly-alpha-methylstyrene (PAMS), and then a GDP layer is overcoated on the surface of
the PAMS mandrel. The GDP shell can be obtained when the PAMS mandrel is depoly-
merized by heating. To prepare good GDP shell, the high degradation yield and low
decomposition temperature of PAMS is important.

Received 17 March 2013; accepted 28 August 2015.


Address correspondence to Zhanwen Zhang, Research Center of Laser Fusion, China Academy
of Engineering Physics, Mianyang 621900, China. E-mail: bjzzw1973@163.com; Junxiao Yang,
School of Material Science and Engineering, Southwest University of Science and Technology,
Mianyang 621010, China. E-mail: yangjunxiao@swust.edu.cn
Color versions of one or more of the figures in the article can be found online at
www.tandfonline.com/lmsb.

1479
1480 L. Ye et al.

It has been shown that the molecular weight of PAMS has a great effect on its decom-
position temperature.[917] The decomposition temperature of PAMS, similar to many
other polymers, decreases with increasing molecular weight (MW).[1117] It has been
shown that the decomposition rate coefficient of PAMS is linear at low molecular weight
but second order at higher molecular weight, implying that the degradation mechanism of
PAMS changes with varying its molecular weight. Reich et al. explained the increase in
the rate at higher MW by enhanced intermolecular hydrogen abstraction, leading to a
chain scission.[16] Another explanation was that longer chains showed a greater tendency
to form tight curves. The resulting bond-angle strain would cause a higher C-C scission
rate.[17] Thus, it could be reasonably concluded that an increase in molecular weight
would lead to an increasing contribution of random scission, which produces a different
structure of the end groups.
The degradation conditions also have an important impact on the thermal degradation
of polymers by altering the degradation mechanism.[1822] The degradation of PAMS in
vacuum was explained by a complex model combining the depolymerization and random
scission.[19] The degradation of PAMS in solutions under pressure occurred by a depo-
lymerization with a low decomposition temperature.[20,21] However, this approach would
be unavailable to use in the preparation of the GDP shells because the hollow structure
could be destroyed by the pressure and some impurities may be introduced by these sol-
vents. It has also been reported that the degradation of PAMS was enhanced in the pres-
ence of peroxides[22] due to increasing the abstraction of hydrogen from the polymer.
However, the enhancement in degradation would be accompanied with the alteration of
the degradation mechanism.
Although the degradation of PAMS under several conditions has been studied,
the degradation of PAMS in an inert atmosphere, as generally used in the prepara-
tion of the GDP shells, has not been studied to our knowledge. Therefore, we inves-
tigated the thermal degradation of PAMS with different molecular weights under
nitrogen atmosphere by thermogravimetric analysis (TGA) and pyrolysis-gas chroma-
tography-mass spectrometry (Py-GC/MS). The effect of molecular weight on the
maximum decomposition rate temperatures, Tmax (the temperature at the maximum
weight loss rate), was studied. The degradation kinetics parameters, such as the acti-
vation energy and the reaction order, were calculated using the Kissinger,[23] the
CoatsRedforn,[24] and Criado methods.[25] According to the degradation kinetics,
the effect of the molecular weight on the degradation mechanisms of PAMS under
nitrogen atmosphere was revealed.

Experimental

Materials
Poly-alpha-methylstyrene was prepared by an anionic polymerization method.[2628] The
monomer, alpha-methylstyrene, was purchased from Aldrich (USA) and was distilled
prior to use. The PAMS was dissolved in toluene and then precipitated in ethanol to
remove lithium hydroxide originating from the synthesis of the PAMS. The produced
powders were dried at 80 C in vacuum for 36 h. Three PAMS samples, designated as
PAMS-1, PAMS-2, PAMS-3, respectively, were used for the degradation measurements.
The gel permeation chromatography (GPC) analysis results of these PAMSs were as fol-
lows: PAMS-1 Mw: 1.13 106, polydispersity index (PDI): 1.55; PAMS-2 Mw: 0.76
106, PDI: 1.92; PAMS-3 Mw: 0.23 106, PDI: 1.04.
Thermal Degradation of PAMS 1481

Apparatus and Measurements


The TGA was carried on a Perkin-Elmer Pyris-1 thermal analyzer instrument (USA) at a
nitrogen atmosphere flow of 20 ml/min, with different heating rates (5 C/min, 10 C/min,
20 C/min, and 30 C/min). The temperature range was from 50 C to 650 C. The Py-GC/
MS experiments was conducted on the Perkin-Elmer Clarus 680 GC and the Perkin-
Elmer Clarus 600 MS spectrometers (USA). Approximately 1 mg samples were placed in
a quartz tube (1 mm internal diameter 25 mm length), which were sealed by some glass
wool. The samples were then loaded in the pyrolysis probe CDS 5200 and placed in the
special inlet in the interface. The pyrolysis temperature was rapidly raised to 500 C with
a heating rate of 10 C/ms and was kept at 500 C for 10 s. The GC oven was programmed
from 60 C to 220 C at a heating rate of 20 C/min. The flow rate of the carrier gas, high
purity helium (99.999%), was set to 10 ml/min. The effluent from the GC spectrometer
entered directly into the source of the MS spectrometer. The total ion current (TIC) and
select ion current (SIC) were plotted as a function of time. The masses were scanned
from m/z 18 to 600. The identifications of the compounds were confirmed by searching
the NIST08 MS library software.

Kinetics Methods
In the kinetics methods employed in this study, the usual symbols involved are summa-
rized as follows:

E: apparent activation energies (kJ/mol)


A: pre-exponential factor (min1), depending only on the nature of the reaction and
independent with the reaction temperature and the concentration of the substance
of the system.
n: apparent reaction order
R: gas constant (8.3136 J/mol/K)
T: absolute temperature (K)
a: fractional weight loss or monomer formation yield
t: reaction time (s)
b: heating rate in thermogravimetric analysis (K/min)
k: rate constant associated with the temperature

The basic assumption of the degradation kinetics is that the Arrhenius equation can
be used to describe the thermal degradation reaction.[29] For a simple degradation reaction
of polymer, its degradation rate da/dt can be given by
da
D kf .a/; (1)
dt
where a D (m0 mt)/m0 (m0 and mt are the initial and the actual weights at time t of the
sample, respectively, in the TGA test) under the assumption that the final weight of
PAMS is zero and the function f(a) depends on the specific decomposition reaction mech-
anism. Here, k can be calculated by the Arrhenius equation

k D Ae E/RT : (2)
1482 L. Ye et al.

Therefore, the degradation rate can be written as follows:

da
D Af .a/e E/RT : (3)
dt

If the heating rate of TGA is set asb D dT


dt , the da/dt in Eq. (3) can be written as

da da dT da
D  Db : (4)
dt dT dt dT

The fractional weight loss, a, thus, is then expressed as a function of the tem-
perature. The combination of these equations, (3) and (4), gives the following
expression:

da A
D f .a/e E/RT (5)
dT b

Both sides of Eq. (5) were integrated and its rearrangement gives

Za ZT
da A
g a D D e E/RT dT (6)
f a b
0 T0

where g(a) is the integral function of conversion degree a. For the different solid
reaction mechanisms, g(a) has different expressions.[30] T0 is the value of T when a
D 0.
Let x D E/RT and then Eq. (6) becomes

Z1
AE
g.a/ D .e x /x2 /dx: (7)
bR
x0

Kissinger Methods[23]
The Kissinger method can be used to calculate the activation energy of the degradation
reaction, adopting the following equation:

 
b AR n1 E
ln. / D ln C ln[n.1 a max / ] ; (8)
T 2 max E RTmax

where Tmax represents the temperature at the maximum degradation rate and amax repre-
sents the conversion degree at the maximum degradation rate. E is calculated from the
slope of the fitted straight line produced from the plot of ln (b/T2max) against 1/Tmax.
Thermal Degradation of PAMS 1483

CoatsRedfern Method[24]
The CoatsRedfern model can be used to calculate the kinetic parameters through the
main thermal degradation region. Compared with the model of Kissinger, the Coats
Redfern model has advantages of simplicity and a shortcut. It deals with the main degra-
dation region of the TGA curves and only requires the TGA data at just one heating rate
to calculate the related reaction order n, activation energy E, and pre-exponential factor,
A.
The CoatsRedfern method adopts the following equations:

    
ln.1 a/ AR 2RT E
nD1 ln D ln 1 (9a)
T2 bE E RT
!   
1 1 a 1 n AR 2RT E
n 6 1 ln D ln 1 : (9b)
T 2 1 n bE E RT

If E/RT >> 1, (1 2RT/E)  1. Thus, both terms on the right of Eqs. (9) can be
regarded as a constant.[31] Then, the determination of the kinetic parameters can be calcu-
lated by the linear fitting of the left part of the above equations dependence on 1/T.
First, we supposed that a certain thermal degradation reaction had a specified reaction
order n and then substituted the supposed n value into above equation. The plot of the left
part of Eq. (9) against 1/T was fitted by computer to calculate the correlation coeffi-
cient. According to the best correlation coefficient, the thermal degradation reaction order
n was determined. Subsequently, the activation energy and the pre-exponential factor can
be calculated from the slope and intercept of the fitted straight line, respectively.

Criado Method[25]
In contrast to the CoatsRedfern method, the Criado method[25] can determine the specific
mechanism in a solid reaction process, such as the thermal decomposition reaction mech-
anism, from dynamic TGA curves for all polymers.
First, Criado et al. defined a new function, Z(a):

da
Z a D dt
PxT; (10)
b

Table 1
Tmax of PAMSs with different molecular weights at different heating rates

Tmax ( C)
Heating rate ( C/min) PAMS-1 PAMS-2 PAMS-3
5 304.87 319.13 324.82
10 313.49 330.80 336.26
20 331.38 346.04 349.15
30 346.07 358.93 362.25
1484 L. Ye et al.

where x D E/RT and P(x) is an approximate expression obtained by the integration of x


against the temperature. Paterson[32] proposed a reasonable relationship between the P(x)
and P(x):

P.x / D xex P.x/; (11)

Figure 1. Pyrolysis products of PAMS-1, PAMS-2, and PAMS-3 at 500 C: (a) GC plot, (b) MS plot
for t D 3.2 min, and (c) MS plot for t D 5.3 min.
Thermal Degradation of PAMS 1485

Figure 2. TGA and DTGA curves of PAMS of different molecular weights at a heating rate of
10 C /min in nitrogen.

where

Z1
P.x/ D .e x /x2 /dx: (12)
x

The combination of Eqs. (3), (7), (10), (11), and (12) leads to the following relation-
ship:

Z.a/ D f .a/g.a/: (13)

Combining Eqs. (5) and (13) results in the following equation:

b da
Z a D ga eE/RT : (14)
A dT

From Eqs. (4), (10), and (11), the following relationship can also be obtained:

da da E E/RT da E E/RT
Z a D x exp.x/P.x/T D e P.x/T D e P.x/; (15)
dT dT RT dT R

where x D E/RT, and

ex x3 C 18x2 C 86x C 96
Px D (16)
x x4 C 20x3 C 120x2 C 240x C 120

When x > 20, Eq. (16) can give an error less than 105%, which is the basis for ana-
lyzing the TGA data. This is an approximate expression of P(x), which was proposed by
Senum and Yang[33] and Flynn.[34] da/dT can be calculated from the related DTGA
(derivative thermogravimetric analysis) curve.
Equation (14) was used to plot the master Z(a) a curve of the different reaction
mechanisms shown in Table 1, and Eq. (15) was used to plot the experimental Z(a) a
1486 L. Ye et al.

Figure 3. TGA and DTGA curves of PAMS of different molecular weights at different heating
rates in nitrogen: (ab) PAMS-1; (cd) PAMS-2; and (ef) PAMS-3.

curve according to the TGA data. The comparison of the master Z(a) a curve with the
experimental Z(a) a curve can easily and accurately predict the reaction mechanism of
the thermal degradation reaction.

Results and Discussion

Py-GC/MS Analysis
The Py-GC/MS was employed to analyze the thermal degradation products arising from
the pyrolysis of PAMS with different molecular weights. The pyrolysis was carried out at
500 C. The GC and MS spectra of the resulting volatilized products from the PAMS-1,
Thermal Degradation of PAMS 1487

Figure 4. Plots of ln(b/T2max) against 1/Tmax using the Kissinger method: (a) PAMS-1;
(b) PAMS-2; and (c) PAMS-3.

PAMS-2, and PAMS-3 are depicted in Fig. 1. For the PAMS-1, PAMS-2, and PAMS-3,
the GC spectrum were almost the same, showing only two peaks, appearing at around
3.2 min and 5.3 min (Fig. 1a). The identifications of the MS spectrum were confirmed by
1488 L. Ye et al.

Table 2
Activation energies of PAMSs calculated by the Kissinger method

Sample E (kJ/mol) Correlation coefficient (R)


PAMS-1 115 0.94
PAMS-2 130 0.98
PAMS-3 143 0.97

searching the NIST08 MS library software. As shown in Fig. 1b, the corresponding MS
spectrum of the product eluting at around 3.2 min presented large ion signals at m/z of 91
and 92, which are the typical ion peaks of toluene (residual solvent in the PAMS) despite
the sample having been dried in vacuum for 36 h. The peak at around 3.2 min of PAMS-
3 was higher than that of PAMS-1 and PAMS-2, indicating that the content of toluene in
the PAMS-3 was the largest. In Fig. 1c, the MS spectrum of the product eluting at around
5.3 min showed ion signals at m/z of 117, 118, and 119, which are the typical ion peaks
of alpha-methylstyrene. There were many other peaks in the MS spectrum of toluene and
alpha-methylstyrene, due to many fragments broken from the toluene and alpha-methyl-
styrene, respectively. From the results of the Py-GC/MS, it is clearly seen that the pyroly-
sis products of PAMS were mainly alpha-methylstyrene and the molecular weights had
no significant effects on the degradation products of PAMS.

Thermogravimetric Analysis
Figure 2 shows the TGA and DTGA curves of PAMS-1, PAMS-2, and PAMS-3 under
nitrogen atmosphere at a heating rate of 10 C/min. One can see that Tmax decreased with
increasing molecular weight, reflecting that the major degradation process of PAMS
shifted to lower temperatures. Moreover, the residual weight of all tested samples was
almost zero, indicating a complete degradation. In particular, in contrast to PAMS-2 and
PAMS-3, Tmax and the temperature at which the residual mass became nearly zero of
PAMS-1 showed a remarkable decrease of 15 C20 C. Such significant reduction of
Tmax and the end of degradation demonstrates the superiority of high molecular weight
PAMS as the mandrel material since a low decomposition temperature is desired.
For the application of PAMS in the preparation of the GDP shells, nearly 100%
monomer yield is also required. To achieve this, the decomposition of PAMS should
obey a depolymerization mechanism. Thus, a crucial question proposed herein is what
the decomposition mechanism of PAMS in nitrogen is. Furthermore, as mentioned above,
an increase in molecular weight of PAMS results in an alteration of the degradation mech-
anism in vacuum, leading to a reduction in monomer yield.[14] Therefore, the effects of
molecular weight on the decomposition mechanism of PAMS is worth investigating in
depth. In this study, the degradation mechanism of PAMS was investigated according to
the data provided by the TGA. The activation energy of degradation for the tested sam-
ples was also calculated from the TGA data.

Kinetic Analysis
In order to carry out the kinetic analysis, TGA was conducted at different heating rates, 5, 10,
20, and 30 C/min. The obtained TGA and DTGA curves are shown in Fig. 3 with Tmax listed
Thermal Degradation of PAMS 1489

in Table 1. For PAMS-1, PAMS-2, and PAMS-3, it is obvious that the Tmax increased with
increasing heating rate, which is attributed to their poor heat conductivity.[29] However, com-
pared with PAMS-1 and PAMS-2, the degradation of PAMS-3 started a little earlier (Fig. 3e
and f), probably due to there being more residual solvent in PAMS-3, which is consistent with
the GC/MS results.
1490 L. Ye et al.

Table 3
Kinetic parameters of PAMS with different molecular weights at the optimum coefficient
obtained using CoatsRedfern method at different heating rates

Heating rate Reaction Activation Pre-exponential Correlation


Sample (b,  C /min) order (n) energy (E, kJ/mol) factor (lnA) coefficient (R)
PAMS-1 5 1 165 36.74 0.9992
10 1 156 34.63 0.9983
20 1 139 30.36 0.9955
30 1 168 35.35 0.9994
Average 1 157 34.27
PAMS-2 5 1 179 39.07 0.9932
10 1 186 39.79 0.9996
20 1 185 38.71 0.9996
30 1 168 34.71 0.9983
Average 1 180 38.07
PAMS-3 5 1 198 42.87 0.9993
10 1 184 39.47 0.9951
20 1 188 39.31 0.9979
30 1 191 38.97 0.9940
Average 1 190 40.15

Kinetics Analysis using the Kissinger Method


The Kissinger method was used to calculate the degradation activation energy of PAMS.
Based on Eq. (8), straight lines were fit to the plots of ln(b/T2max) against 1/Tmax
(Fig. 4). The apparent activation energy E can be calculated from the slopes of the
straight lines (Table 2). Although the data would better fit curved lines, as can be seen,
the correlation coefficients of the various fitted straight lines were higher than 90%, indi-
cating that the fit was reliable. The calculated activation energies of the thermal degrada-
tion for PAMS-1, PAMS-2, and PAMS-3 were 115 kJ/mol, 130 kJ/mol, and 143 kJ/mol,
respectively, indicating a facilitated decomposition with increasing molecular weight.
This is consistent with the reduction in the Tmax with increasing molecular weight, show-
ing the applicability of the Kissinger method.

Kinetics Analysis using the CoatsRedfern Method


The CoatsRedfern method presents advantages in providing not only the activation
energy, but also the kinetic parameters, such as the reaction order and the pre-exponential
factor, which are related to the degradation mechanism. To calculate the kinetic parame-
ters mentioned above, only the Tmax at one heating rate is required. Figure 5 shows the
linear fitting process for the different molecular weights of PAMS for various preselected
reaction orders at 10 C/min heating rate, according to Eq. (6). When the value of n was
1, the fitting was the best. The best-fitted results, including the reaction orders and the
activation energies of PAMS-1, PAMS-2, and PAMS-3 for the various heating rates, are
summarized in Table 3. The reaction orders of all samples were close to 1, indicating a
simple degradation mechanism in nitrogen for PAMS, independent of the molecular
Thermal Degradation of PAMS 1491

Figure 6. Determination of the thermal degradation reaction mechanism by plots of Z(a) vs. a using
the Criado model: (a) PAMS-1; (b) PAMS-2; and (c) PAMS-3.

weight and the heating rate. This first-order reaction mechanism should be linked with the
dominant depolymerization mechanism, with random scission being negligible. More-
over, in agreement with the result obtained by Kissinger method, it was also found that
the activation energy decreased with increasing molecular weight (Table 3).
1492 L. Ye et al.

Table 4
Expressions of g(a) for the most frequently used reaction mechanisms of solid state
processes[30]

Mechanism g(a) Solid state process description


Sigmoidal function
A2 [ln(1 a)] 1/2
Nucleation and growth: Avrami Eq. (1)
A3 [ln(1 a)]1/3 Nucleation and growth: Avrami Eq. (2)
A4 [ln(1 a)]1/4 Nucleation and growth: Avrami Eq. (3)
Deceleration function
R2 [1 (1 a)1/2] Phase boundary controlled reaction: contraction area
R3 [1 (1 a)1/3] Phase boundary controlled reaction: contraction
volume
D1 a2 One-D diffusion
D2 (1 a)ln(1 a) C a Two-D diffusion
D3 [1 (1 a)1/3] 2 Three-D diffusion: Jander equation
D4 (1 2/3a) (1 a)2/3 Three-D diffusion: GinstlingBrounshtein equation
F1 ln(1 a) Random nucleation having onenucleus on
individual particle
F2 1/(1 a) Random nucleation having twonucleus on
individual particle

Estimation of Mechanism using the Criado Method


The kinetic parameters at 10 C/min heating rate, obtained by the CoatsRedfern method
(Table 3), were substituted into Eqs. (14) and (15). The Z(a) a master curve was then
plotted using Eq. (14) according to the different reaction mechanisms, g(a), shown in
Table 4, where the experimental TGA data were still from the TGA curves recorded at
10 C/min heating rates. The corresponding Z(a) a experimental curves were plotted
using Eq. (15) in Fig. 6, which also shows the Z(a) a master and experimental curves
for the different molecular weights of PAMS. One can clearly see that the experimental
curves of the three samples of PAMS nearly overlap the master curve Z(F1), indicating
that the thermal degradation of the three samples of PAMS corresponded to the F1 reaction
mechanism (random nucleation having one nucleus on an individual particle). This mecha-
nism should thus be linked with a depolymerization mechanism where each polymer chain
has only one active degradation site.[3537] In contrast, a random scission mechanism, in
principle, leads to the production of two or even more degradation sites. Moreover, accord-
ing to g(a) D ln(1 a) and Eq. (6), we can obtain that f(a) D 1 a. This result indi-
cated that the F1 mechanism was characteristic of a first-order degradation reaction,[38]
which is in accordance with the results obtained by the CoatsRedfern method.
The above results indicated that the degradation of PAMS in nitrogen obeys a depo-
lymerization model independent of the molecular weight. As mentioned in the literature,[18
22]
under other conditions, such as in vacuum or a closed system, the degradation mechanism
of PAMS follows a complex model and an increase in molecular weight generally leads to
an increase in the contribution of random scission. However, the reason for this unique
behavior is not clear at the present stage and further studies are now on-going in our group.
Thermal Degradation of PAMS 1493

Conclusions
The thermal degradation behavior of PAMS with different molecular weights in nitrogen
was analyzed in this work. The TGA results showed that Tmax of PAMS decreased with
increasing molecular weight. The calculation results from the Kissinger and Coats
Redfern methods showed that the activation energy of PAMS degradation also decreased
with increasing molecular weight, which explained well the decrease in the decomposi-
tion temperature. Both the CoatsRedfern and Criado methods revealed a first-order
decomposition reaction model for the PAMS with different molecular weights in nitro-
gen. This demonstrated that the depolymerization mechanism of PAMS was maintained
at high MW when the pyrolysis was conducted in nitrogen. This finding is suggested to
provide a new route to enhance the degradation efficiency of PAMS, as desired for its use
in the preparation of GDP shells.

Funding
The financial support from the National High Technology Development Project Founda-
tion of China and the China Academy of Engineering Physics (2014B0302052) is grate-
fully acknowledged.

References
1. Grassie, N.; Scott, G. Polymer Degradation and Stabilisation; Cambridge University Press:
New York, 1985.
2. Zhong, S.Y.; Xu, Q.W.; Wang, G.S. Polymer Degradation and Stability; Chemistry Industry
Press: Beijing, 2002, p. 28.
3. Fearon, E.M.; Letts, S.A.; Allison, L.M.; Cook, R.C. Adapting the decomposable mandrel tech-
nique to build specialty ICF target. Fusion Technol. 1997, 31, 406410.
4. McQuillan, B.W.; Nikroo, A.; Steinman, D.A.; Elsner, F.H.; Czechowicz, D.G.; Hoppe, M.L.;
Sixtus, M.; Miller, W.J. The PAMS/GDP process for production of ICF target mandrels. Fusion
Technol. 1997, 3, 381384.
5. Takagi, M.; Cook, R.; McQuillan, B.; Gibson, J.; Paguio, S. Investigation of large poly
(a-methylstyrene) mandrels for high gain designs using microencapsulation. Fusion Technol.
2004, 45, 171175.
6. Zhang, Z.W.; Qi, X.B.; Tang, Y.J.; Li, B.; Wang, C.Y.; Huang, Y. Theoretical simulation of
PVA coating using drop-tower technique. High Power Laser Part. Beams. 2006, 11, 1837
1840.
7. Bhandarkar, S.; Letts, S.A.; Buckley, S.; Alford, C.; Lindsey, E.; Hughes, J.; Youngblood, K.P.;
Moreno, K.; Xu, H.; Huang, H.; Nikroo, A. Removal of the from beryllium sputter coated cap-
sules for NIF targets. Fusion Technol. 2007, 51, 564571.
8. Zhang, Z.W. Foundational Study of Fabrication Technology of Double-Shell Targets in Inertial
Confinement Fusion; China Academy of Engineering Physics: Mianyang, 2007, p. 117119.
9. Jellinek, H.H.G. Thermal degradation of polystyrene and polyethylene. Part III. J. Polym. Sci.
1949, 4, 2528.
10. Brown, D.W.; Wall, L.A. Pyrolysis of poly-a-methylstyrene. J. Phys. Chem. 1958, 62, 848852.
11. Ferriol, M.; Gentilhomme, A.; Cochez, M.; Oget, N.; Mieloszynski, J.L. Thermal degradation
of poly (methyl methacrylate) (PMMA): Modeling of DTG and TG curves. Polym. Degrad.
Stab. 2003, 79, 271281.
12. Manring, L.E.; Sogah, D.Y.; Cohen, G.M. Thermal degradation of poly(methyl methacrylate).
3. Polymer with head-to-head linkages. Macromolecules 1989, 22, 46524654.
1494 L. Ye et al.

13. Kashiwagi, T.; Inaba, A.; Brown, J.E.; Hatada, K.; Kitayama, T.; Masuda, E. Effects of weak
linkages on the thermal and oxidative degradation of poly(methyl methacrylates). Macromole-
cules 1986, 19, 21602168.
14. Malhotra, S.L.; Baillet, C.; Minh, L.Y.; Blanchard, L.P. Thermal decomposition of poly-a-
methylstyrene. J. Macromol. Sci. Part A. Chem. 1978, 12, 129147.
15. Madras, G.; Chung, G.Y.; Smith, J.M.; McCoy, B.J. Molecular weight effect on the dynamics
of polystyrene degradation. Ind. Eng. Chem. Res. 1997, 36, 20192024.
16. Reich, L.; Stivala, S.S. Elements of Polymer Degradation; McGraw-Hill: New York, 1971,
p. 186.
17. Vollhardt, K.P.C.; Schore, N.E. In Organic Chemistry, 2nd ed.; Freeman: New York, 1994,
-p. 111.
18. Straus, S.; Madorsky, S.L. Pyrolysis of styrene, acrylate, and isoprene polymers in a vacuum. J.
Res. Nat. Bur. Stand. 1953, 50, 165174.
19. Jellinek, H.H.G.; Kachi, H. Thermal degradation of poly(a-methylstyrene) in a closed system. J.
Polym. Sci. C 1968, 23, 97108.
20. Bywater, S.; Black, P.E. Thermal depolymerization of polymethyl methacrylate and poly-
a-methylstyrene in solution in various solvents. J. Phys. Chem. 1965, 69, 29672970.
21. Madras, G.; Smith, J.M.; McCoy, B.J. Thermal degradation of poly(a-methylstyrene) in solu-
tion. Polym. Degrad. Stab. 1996, 52, 349358.
22. Sterling, W.J.; Kim, Y.-C.; McCoy, B.J. Peroxide enhancement of poly(a-methyl styrene) ther-
mal degradation. Ind. Eng. Chem. Res. 2001, 40, 18111821.
23. Kissinger, H.E. Reaction kinetics in differential thermal analysis. Anal. Chem. 1957, 29,
17021706.
24. Coats, A.W.; Redfern, J.P. Kinetic parameters from thermodynamic data. Nature 1964, 201,
6869.
25. Criado, J.M.; Malek, J.; Ortega, A. Applicability of the master plots in kinetic analysis of non-
isothermal data. Thermochim. Acta 1989, 147, 377385.
26. Tian, G.H.; Ye, L.N.; Huang, Y.W.; Cao, K.; Yang, J.X. Controllable preparation of poly-
a-methylstyrene polymer. Atom. Energy Sci. Technol. 2012, 46, 903906.
27. Bitsenko, M.I.; Podolsky, A.F.; Khvostik, G.M.; Sokolov, V.N. Polymerization of a-methyl-
styrene by n-butyllithium in tetrahydrofuran. J. Polym. Sci. A 1972, 10, 32053215.
28. Worsfold, D.J.; Bywater, S. Anionic polymerization of a-methylstyrene. J. Polym. Sci. 1957,
26, 299304.
29. Liu, Z.H. Introduction of Thermal Analysis; Chemistry Industry Press: Beijing, 1991, p. 9.
30. Nu~nez, L.; Fraga, F.; Nu~nez, M.R.; Villanueva, M. Thermogravimetric study of the decomposi-
tion process of the system BADGE (n D 0)/1,2 DCH. Polymer 2000, 41, 46354641.
31. Hu, R.Z.; Shi, Q.Z. Thermal Analysis Kinetics; Science Press: Beijing, 2001, p. 4749.
32. Paterson, W.L. Computation of the exponential trap population integral of glow curve theory. J.
Comput. Phys. 1971, 7, 187190.
33. Senum, G.I.; Yang, R.T. Rational approximations of the integral of the Arrhenius function. J.
Therm. Anal. Calorim. 1977, 11, 445449.
34. Flynn, J.H. The temperature integral: Its use and abuse. Thermochim. Acta 1997, 300, 8392.

35. Satava, 
V.; Skvara, F. Mechanism and kinetics of the decomposition of solids by a thermogravi-
metric method. J. Am. Cer. Soc. 1969, 52, 591595.
 ak, J.; Berggren, G. Study of the kinetics of the mechanism of solid-state reactions at
36. Sest
increasing temperatures. Thermochim. Acta 1971, 3, 112.
37. Ma, S.; Hill, J.O.; Heng, S. A kinetic analysis of the pyrolysis of some Australian coals by non-
isothermal thermogravimetry. J. Therm. Anal. 1991, 37, 11611177.
38. Alshehri, S.M.; Monshi, M.A.S.; Abd El-Salam, N.M.; Mahfouz, R.M. Kinetics of the thermal
decomposition of g-irradiated cobaltous acetate. Thermochim. Acta 2000, 363, 6170.

You might also like