06.04.2013 Views

저작자표시-비영리-변경금지 2.0 대한민국 이용자는 아래의 조건을 ...

저작자표시-비영리-변경금지 2.0 대한민국 이용자는 아래의 조건을 ...

저작자표시-비영리-변경금지 2.0 대한민국 이용자는 아래의 조건을 ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>저작자표시</strong>-<strong>비영리</strong>-<strong>변경금지</strong> <strong>2.0</strong> <strong>대한민국</strong><br />

<strong>이용자는</strong> <strong>아래의</strong> <strong>조건을</strong> 따르는 경우에 한하여 자유롭게<br />

l 이 저작물을 복제, 배포, 전송, 전시, 공연 및 방송할 수 있습니다.<br />

다음과 같은 <strong>조건을</strong> 따라야 합니다:<br />

<strong>저작자표시</strong>. 귀하는 원저작자를 표시하여야 합니다.<br />

<strong>비영리</strong>. 귀하는 이 저작물을 영리 목적으로 이용할 수 없습니다.<br />

<strong>변경금지</strong>. 귀하는 이 저작물을 개작, 변형 또는 가공할 수 없습니다.<br />

l 귀하는, 이 저작물의 재이용이나 배포의 경우, 이 저작물에 적용된 이용허락조건<br />

을 명확하게 나타내어야 합니다.<br />

l 저작권자로부터 별도의 허가를 받으면 이러한 조건들은 적용되지 않습니다.<br />

저작권법에 따른 이용자의 권리는 위의 내용에 의하여 영향을 받지 않습니다.<br />

이것은 이용허락규약(Legal Code)을 이해하기 쉽게 요약한 것입니다.<br />

Disclaimer


工 學 碩 士 學 位 論 文<br />

Poly(butylene succinate)(PBS)/탄소나노튜브,<br />

탄소나노섬유 나노복합체 발포체의 제조와 특성 분석<br />

Preparation and Characterization of Poly(butylene<br />

succinate)(PBS)/Carbon Nanotube, Carbon<br />

Nanofiber Nanocomposite Foams<br />

2009年 2月<br />

仁荷大學校 大學院<br />

高分子工學科<br />

李 錫 仁


工 學 碩 士 學 位 論 文<br />

Poly(butylene succinate)(PBS)/탄소나노튜브,<br />

탄소나노섬유 나노복합체 발포체의 제조와 특성 분석<br />

Preparation and Characterization of Poly(butylene<br />

succinate)(PBS)/Carbon Nanotube, Carbon<br />

Nanofiber Nanocomposite Foams<br />

2009年 2月<br />

指導敎授 陳 仁 柱<br />

이 論文을 碩士學位 論文으로 提出함<br />

仁荷大學校 大學院<br />

高分子工學科<br />

李 錫 仁


이 論文을 李 錫 仁의 碩士學位論文으로 認定함.<br />

2009年 2月<br />

主審<br />

副審<br />

委員


Abstract<br />

Recently, due to the environmental problems caused by synthetic<br />

non-degradable plastic waste, biodegradable polymers have received<br />

extensive attention as the potential candidate to replace the non-<br />

degradable counterparts. One of the promising synthetic aliphatic<br />

polyesters is poly(butylene succinate) (PBS), which has a high melting<br />

temperature (115 o C), good biodegradability by microorganism, and<br />

outstanding processability. However, it is known that its tensile<br />

properties, gas barrier properties and melt viscosity for further<br />

processing are not quite desirable in some end-use applications.<br />

In the polymer nanocomposite foam, the nanoscale filler distributed<br />

in the polymer matrix may act as the nucleation site during the foaming<br />

process, inducing smaller cell size and increased cell density. In<br />

addition, the presence of well-dispersed fillers can improve mechanical<br />

and physical properties, and the heat distortion temperature of polymer<br />

foams. The cell morphology, which depends strongly on the interaction<br />

with the filler, is expected to affect such characteristic properties as<br />

weight, specific strength, heat conductivity and isolation to sound.<br />

The purpose of this study is to develop the biodegradable polymer<br />

nanocomposite foam having high blowing ratio and closed-cell<br />

i


structure. Multi-walled carbon nanotube(MWNT) and carbon<br />

nanofiber(CNF) were used as the nano-sized filler for PBS. After<br />

establishing the optimum condition for producing the polymer<br />

nanocomposite foam, we prepared the PBS nanocomposite foams<br />

using a chemical blowing agent. Physico-chemical properties of the<br />

PBS nanocomposite foams were characterized.<br />

Keyword : Poly(butylene succinate)(PBS), Carbon Nanotube,<br />

Carbon Nanofiber, Nanocomposite Foam, Chemical Blowing Agent<br />

ii


국 문 요 약<br />

최근 비분해성 플라스틱 폐기물로 인한 환경문제가 대두되면서<br />

비분해성 고분자를 대체하기 위한 일환으로 생분해성 고분자가 각<br />

광을 받아 왔다. 그 중 가장 유망한 합성 지방족 폴리에스터 계열<br />

중의 하나인 Poly(butylene succinate)(PBS)는 115 o C 근처의 높은<br />

융점을 가지며 미생물에 의한 생분해도가 높고, 가공성이 훌륭하다.<br />

그러나 인장 성질, 가스 차단성, 용융점도가 낮아 최종 응용적인 측<br />

면에서 제한을 받아 왔다.<br />

고분자 나노복합체 발포체에서 고분자 매트릭스 내에 분산되어<br />

있는 나노사이즈의 충전제는 발포과정에서 기핵제로서의 역할을 할<br />

수 있으며 더 작은 셀의 크기와 셀 밀도의 증가를 유도시킬 수 있<br />

다. 더욱이 잘 분산된 나노사이즈의 충전제로 인해 고분자 발포체의<br />

기계적, 물리적 성질, 열 변형 온도 등의 물성을 향상시킬 수 있다.<br />

충전제와의 상호작용에 강하게 의존하는 셀의 형태는 발포체의 무<br />

게(밀도), 강도, 열 전도도, 차음성과 같은 특성에 영향을 준다.<br />

본 실험의 목적은 고배율의 밀폐형 구조를 가지는 생분해성 고분<br />

자 나노복합체 발포체를 개발하는 것이다. 이에 나노사이즈의 충전<br />

제로서 탄소나노튜브와 탄소나노섬유를 이용하였다. 발포체 제조를<br />

위한 고분자 나노복합체의 최적 <strong>조건을</strong> 확립한 후 화학발포제를 이<br />

용하여 고분자 나노복합체 발포체를 제조하고, 그들의 특성을 분석<br />

하였다.<br />

iii


Keyword: Poly(butylene succinate)(PBS), 탄소나노튜브,<br />

탄소나노섬유, 나노복합체 발포체, 화학발포제<br />

iv


CONTENTS<br />

Abstract ………………………………………………………………… i<br />

국문요약 ……………………………………………………………… iii<br />

CONTENTS …………………………………………………………… v<br />

LIST OF TABLES …………………………………………………… vii<br />

LIST OF FIGURES …………………………………………………… vii<br />

SECTION I. Processing of Poly(butylenes succinate)(PBS)/Carbon<br />

Nanotube Nanocomposite Foams<br />

1. Introduction…………………………………………………………… 1<br />

2. Experimental………………………………………………………… 6<br />

2. 1. Materials………………………………………………………… 6<br />

2. 2. Preparation of polymer nanocomposite……………………… 6<br />

2. 3. Foaming of nanocomposite…………………………………… 8<br />

2. 4. Characterization of nanocomposites and their foams……… 8<br />

3. Results and discussion…………………………………………… 11<br />

3. 1. PBS/MWNTs nanocomposites……………………………… 11<br />

3. 2. Nanocomposite morphology and dispersion of MWNTs… 11<br />

3. 3. Thermal behavior…………………………………………… 12<br />

3. 4. Mechanical testing…………………………………………… 14<br />

3. 5. Foam behavior and cell geometry………………………… 16<br />

4. Conclusions………………………………………………………… 19<br />

5. References………………………………………………………… 21<br />

v


SECTION II. Preparation and Characterization of PBS/Carbon<br />

Nanofiber Nanocomposite Foams<br />

1. Introduction………………………………………………………… 26<br />

2. Experimental………………………………………………………… 29<br />

2. 1. Materials……………………………………………………… 29<br />

2. 2. Nanocomposite preparation………………………………… 29<br />

2. 3. Foaming……………………………………………………… 31<br />

2. 4. Characterization of nanocomposites and their foams…… 31<br />

3. Results and discussion…………………………………………… 33<br />

4. Conclusions………………………………………………………… 39<br />

5. References………………………………………………………… 40<br />

vi


SECTION I.<br />

LIST OF TABLES<br />

Table 1 Morphological parameters of PBS/2 wt% MWNTs<br />

nanocomposite foam obtained by the SOAM method (blowing time = 5<br />

min) …………………………………………………………………… 42<br />

SECTION II.<br />

Table 1 Thermal stability for PBS/CNFs nanocomposites……… 43<br />

Table 2 Temperature and CBA content dependence of morphological<br />

parameters …………………………………………………………… 44<br />

SECTION I.<br />

LIST OF FIGURES<br />

Figure 1 FT-Raman spectra of the PBS nanocomposites with 2 wt%<br />

content prepared by three different methods: (a) MWNT (b) solution<br />

blending (c) melt mixing and (d) SOAM method …………………… 45<br />

Figure 2 SEM images of the PBS nanocomposites with 2 wt% content<br />

prepared by three different methods: (a) solution blending (b) melt<br />

mixing (c) SOAM method, and (d) TEM micrograph of the 2 wt%<br />

PBS/MWNTs nanocomposites prepared by SOAM method … 46<br />

vii


Figure 3 TGA thermograms of the PBS nanocomposites containing<br />

various MWNTs content prepared by the SOAM method ………… 47<br />

Figure 4 DSC traces of the PBS nanocomposites containing various<br />

MWNTs content prepared by the SOAM method: (a) melting and (b)<br />

crystalline behaviors ………………………………………………… 48<br />

Figure 5 Tensile strength and elongation at break curves of the PBS<br />

nanocomposites with various MWNTs content prepared by three<br />

different methods …………………………………………………… 49<br />

Figure 6 Optical micrographs of PBS nanocomposite foams with 2<br />

wt% MWNTs content prepared by the SOAM method as a function of<br />

blowing temperature and CBA content. The blowing time was fixed at 5<br />

min ……………………………………………………………………… 50<br />

Figure 7 Cell size distribution of PBS nanocomposite foams with 2<br />

wt% MWNTs content prepared by the SOAM method at 150 o C: (a) 4<br />

(b) 6 and (c) 8 phr. The blowing time was fixed at 5 min ………… 51<br />

Figure 8 Blowing ratio of PBS nanocomposite foams with 2 wt%<br />

MWNTs content prepared by the SOAM method as a function of<br />

blowing time and temperature and CBA content: (a) 4 (b) 6 and (c) 8<br />

phr ……………………………………………………………………. 52<br />

viii


SECTION II.<br />

Figure 1 TGA thermograms of PBS/CNFs nanocomposites by means<br />

of different preparation methods ……………………………………… 53<br />

Figure 2 The Effect of CNFs content and different preparation<br />

methods on the (a) tensile strength and (b) elongation at break …… 54<br />

Figure 3 SEM images of PBS/2 wt% CNFs nanocomposite prepared<br />

by (a) solution blending (b) melt mixing and (c) SOAM method …… 55<br />

Figure 4 The Effect of blowing time and temperature on the cross-<br />

sectional cell morphology of PBS/2 wt% CNFs nanocomposite foams<br />

measured by the optical microscope. CBA content was fixed at 4 phr<br />

………………………………………………………………………… 56<br />

Figure 5 Blowing temperature and time dependence on the blowing<br />

ratio of PBS/2 wt% CNFs nanocomposite foams with different CBA<br />

loading: (a) 4 (b) 6 and (c) 8 phr ……………………………………… 57<br />

Figure 6 Cell size distribution of the PBS/2 wt% CNFs nanocomposite<br />

foams as a function of the blowing temperature: (a) 150 (b) 160 and (c)<br />

170 o C. The blowing time and CBA content were fixed for 5 min and at<br />

4 phr, respectively …………………………………………………… 58<br />

ix


SECTION I. Processing of Poly(butylenes succinate)(PBS)/Carbon<br />

Nanotube Nanocomposite Foams<br />

1. Introduction<br />

Over the past decade, numerous nano-scaled reinforcing fillers have<br />

been investigated. Among them, study on polymeric composites with<br />

carbon nanotubes (CNTs) has actively carried out due to the unique<br />

properties of the composites. [1-4] The extraordinary properties of CNTs<br />

such as large aspect ratio, low density, high flexibility, modulus and<br />

strength have spurred much research. Especially, it was revealed that<br />

polymer-CNTs nanocomposites exhibit dramatically enhanced thermal,<br />

mechanical and electrical properties as compared with those of pristine<br />

polymer. [5-7] This has led to the utilization of polymer-CNTs<br />

nanocomposites in numerous application fields. [1,8] However, the<br />

following two prior conditions should be established in order to realize<br />

desired performance: (i) a homogeneous dispersion of the CNTs<br />

throughout the polymer matrix and (ii) strong polymer-CNTs interfacial<br />

adhesion. As in most fiber or particle reinforced composite materials, in<br />

order to achieve the most efficient enhancement of properties, the<br />

reinforcement phase should be uniformly dispersed in the polymer<br />

1


matrix. However, CNTs are strongly affected by van der Waals force<br />

due to their small size and large surface area. These forces give rise to<br />

the formation of aggregates, which in turn, make dispersion of CNTs in<br />

the polymers difficult. In order to attain an efficient dispersion of CNTs<br />

by the physical approach, ultrasound can be applied to a dispersion of<br />

CNTs. [9] In addition, chemical methods such as covalent attachment<br />

involving the introduction of a chemical group to the nanotube wall,<br />

and non-covalent adsorption methods have been employed in order to<br />

provide chemical compatibility with the polymers, and dissolution of<br />

CNTs in organic solvents. [10-12]<br />

Generally, the following three methods for the preparation of<br />

polymer-CNTs nanocomposite are employed to promote the dispersion<br />

of controlled CNTs throughout the polymer matrix: (i) solution blending,<br />

a method for precipitating or casting a polymer-CNTs solution<br />

dissolved in a solvent, [13,14] (ii) melt mixing, which requires high<br />

temperature and shear force according to the polymer species in the<br />

mixer without a solvent and (iii) in-situ polymerization, a method to<br />

polymerize the monomer in the presence of the dispersed-CNTs. This<br />

last approach has outstanding potential which can include covalent<br />

bonding between the polymer matrix and modified-CNTs. [16,17]<br />

2


A nanoscale filler dispersed in a polymer matrix can act as a<br />

nucleation agent during the foaming process, resulting decreased cell<br />

size and increased cell density. [18,19] Also, the cell morphology, which<br />

depends strongly on the interaction with the reinforced-filler, is related<br />

with the properties such as lightweight, high specific strength, poor<br />

heat conductivity, and good isolation to sound. And, several groups<br />

have developed physical and chemical foaming methods and<br />

investigated the properties of the final foams. Nam et al. prepared<br />

PP/clay nanocomposite foams by using a supercritical CO2 in the<br />

autoclave. Their products exhibited the homogeneity in cell size of 30<br />

~ 120 μm, the high cell density of about 10 7 ~ 10 8 cell/mL, and the high<br />

cell wall thickness of 5 ~ 15 μm. The clay particles also induced the<br />

improvement of the strength of foam in mechanical properties. Di et al.<br />

reported PLA/organoclay nanocomposite foams formed by using the<br />

mixture of CO2 and N2. The improvements of viscous and elastic<br />

properties for PLA/clay nanocomposites allowed them to produce PLA<br />

foams with much smaller cell size and higher cell density than those of<br />

pure PLA even at low clay content. With supercritical CO2 via the batch<br />

foaming process, CNFs and CNTs have been used to prepare PS<br />

nanocomposite foams by Shen et al. They compared the nucleation<br />

3


efficiency of different particles, suggesting that CNFs exhibited an<br />

excellent nucleation effect due to its food dispersion in the polymer<br />

matrix, as well as the favorable wettability and surface curvature in that<br />

foaming process. On the contrary, using a chemical blowing agent,<br />

Guan et al. manufactured the poly(propylene carbonate) (PPC) foams<br />

with a blowing ratio of 16. The decomposition behavior of the blowing<br />

agent was investigated, and the effect of blowing agent content and<br />

[20-24, 36]<br />

the foaming condition on the foaming of PPC studied.<br />

Recently, synthetic petroleum-based polymers have caused serious<br />

environmental problems due to their limited degradation. For this<br />

reason, biodegradable polymers have drawn substantial attention<br />

owing to environmental concerns. Among these polymers,<br />

biodegradable aliphatic polyester types, i.e., poly(ε-caprolactone)<br />

(PCL), poly(lactic acid) (PLA), poly(3-hydroxybutyrae) (PHB) and<br />

poly(butylene succinate) (PBS), have currently been the focus of much<br />

attention. [25-30] However, they lack price competitiveness and a<br />

practically acceptable level of processability, and their usage is due to<br />

relatively poor physical properties in particular. Therefore, the<br />

incorporation of nanoscale fillers into biodegradable aliphatic<br />

polyesters has been studied by many research groups with the aim of<br />

4


enhancing the physical properties of the polyesters. [31-35] Poly(butylene<br />

succinate) (PBS) is a biodegradable aliphatic polyester resin<br />

synthesized by a condensation reaction between 1, 4-butandiol and<br />

succinic acid, which has a high fusion temperature (115 o C),<br />

remarkable biodegradation, and outstanding processability by means<br />

of the extruder used for polyolefins. [26,29] However, it is known that its<br />

tensile property, gas barrier properties, melt viscosity for further<br />

processing and so on are deficient in various end-use applications. [27]<br />

In this study, three different nanocomposite fabrication approaches<br />

such as solution blending, melt mixing and the SOAM method were<br />

applied to compare the degree of dispersion of MWNTs in the PBS<br />

domain. Herein, we have mainly focused on the SOAM method<br />

because it can achieve a much better MWNTs dispersion by melt-<br />

mixing the nanocomposite prepared by solution blending, under strong<br />

shear force. In order to verify the reinforcing effect of MWNTs as<br />

compared to neat PBS, as well as to establish the optimal conditions<br />

for expanding nanocomposites, the thermal and mechanical properties<br />

were investigated. We obtained PBS nanocomposite foams with a<br />

closed-cell structure using a chemical blowing agent and characterized<br />

the foam behavior.<br />

5


2. 1. Materials<br />

2. Experimental<br />

The Poly(butylene succinate)(PBS) used in this study was Enpol<br />

G4560 (MI = 1.5 g/10 min) from Ire Chemical LTD., Korea. MWNTs<br />

(HollowCNT 75L, purity of higher than 95%, inner/outer diameter<br />

distribution 30~50 / 60~80 nm) were purchased from Nanokarbon Co.,<br />

LTD., Korea. A high purity treatment using HCl and a density treatment<br />

were applied according to the manufacturer’s instructions. Chloroform<br />

was used as a solvent in the solution blending step and was supplied<br />

by Dongyang chemical, Korea. Cellecom ACP-4 acted as a chemical<br />

blowing agent (CBA) and was obtained by Kumyang Co., Korea. All<br />

materials were used without further purification.<br />

2. 2. Preparation of polymer nanocomposites<br />

In this work, three approaches were employed to evaluate the<br />

dispersion of MWNTs in the PBS matrix: solution blending, melt mixing,<br />

and the SOAM method (solution blending and subsequent melt mixing),<br />

respectively. PBS pellets and MWNTs were dried in a vacuum oven at<br />

60 o C for 24 h prior to being used. First, in the case of solution<br />

6


lending, PBS pellets were dissolved in chloroform at 40 o C and stirred<br />

for 5 h to obtain a complete polymer solution. At the same time,<br />

MWNTs were dispersed in chloroform at ambient temperature for 3 h<br />

and then sonicated for 2 h at 90 W by a ultrasonic generator (Kodo<br />

Technical Research Co., Japan). Then, the MWNTs-dispersed solution<br />

was added to PBS solution and then mixed under strong stirring at 40<br />

o C for 24 h to induce a nanocomposite with a uniform dispersion of<br />

MWNTs. After the removal of chloroform under the rotary hood, the<br />

nanocomposite film was dried under a vacuum at 60 o C for 24 h.<br />

Second, melt mixing was performed by a torque rheometer (Haake<br />

Rheometer 90, Germany) at 120 o C for 15 min with a rotor speed of 60<br />

rpm. After the PBS pellets were melted for 3 min, MWNTs having<br />

different contents were gradually added. Third, some nanocomposite<br />

films prepared by solution blending were melt-mixed for 10 min inside<br />

a twin screw mixer under the aforementioned conditions, The SOAM<br />

method exhibited higher torque than that of simple melt mixing at an<br />

early state. The mass fraction of MWNTs loaded to the polymer matrix<br />

was varied from 0.1 to 3 wt%. The sample codes were named, for<br />

example, PBS2S, PBS2M, and PBS2SM, where the numbers<br />

7


epresent the mass fraction of MWNTs, and the ‘S’ stands for solution<br />

blending, ‘M’ for melt mixing, and ‘SM’ for the SOAM method.<br />

2. 3. Foaming of nanocomposites<br />

Nanocomposite films fabricated by solution blending and having a<br />

uniform size (25 mm width × 50 mm long) were put into a Haake<br />

rheometer and mixed into a molten state at 120 o C for 10 min with a<br />

screw speed of 60 rpm. Afterwards, ACP-4 with different contents was<br />

slowly added to the mixer and blended for a further 10 min. The<br />

resulting products were prepared by compression molding at 135 o C<br />

for 8 min into strips of 12 mm × 60 mm × 3 mm. Each strip was<br />

again cut into samples of 12 mm × 20 mm, followed by foaming at<br />

150, 160, and 170 o C for 5, 10, 20, and 30 min in a forced convection<br />

oven, respectively.<br />

2. 4. Characterization of nanocomposites and their foams<br />

FT-Raman spectra were acquired to evaluate the formation of MWNT-<br />

reinforced PBS nanocomposites using a RFS 100/S (Bruker,<br />

Germany) with a spectral resolution of 4 cm -1 . The excitation length<br />

was the 1064 nm line of a 60 mW Nd-YAG laser. The cryo-fractured<br />

8


surface morphology of the nanocomposites was observed by field<br />

emission scanning electron microscopy using a S-4300 (Hitachi,<br />

Japan). The samples were freeze-fractured in liquid nitrogen and<br />

sputter-coated with Pt. TEM observations were performed on a JEM<br />

2100F (JEOL, Japan) operated at an accelerated voltage of 200 kV to<br />

examine the dispersion of MWNTs embedded in the PBS matrix.<br />

Ultrathin sections with a thickness of 50nm were microtomed using a<br />

MTX (RMC, Japan) with a diamond knife. A thermal analysis was<br />

conducted with a modulated DSC 2920 (TA Instruments, USA) in a<br />

nitrogen atmosphere at a heating rate of 10 o C/min. To erase the<br />

thermal history of the nanocomposite, a second heating scan was<br />

applied at temperatures ranging from 30 o C to 170 o C. Thermal stability<br />

was investigated out on a TGA Q50 (TA Instruments, USA). All<br />

samples were heated to 700 o C under air at a heating rate of 20 o C/min.<br />

Mechanical properties of the nanocomposites films under uniaxial<br />

elongation at room temperature were measured using a UTM<br />

(Hounsfield Test Equipment, UK). Each sample was provided as a<br />

dog-bone type and the resulting value was an average of at least<br />

seven measurements.<br />

9


The foam density was measured by a gas pycnometer (Accupyc 1330,<br />

Micrometrics Co., USA). The blowing ratio was calculated using the<br />

following equation:<br />

3<br />

the density of neat PBS (1.268 g/cm )<br />

ratio =<br />

the density of PBS/MWNTs nanocomposite<br />

foam (g/cm<br />

Blowing 3<br />

The cross-sectional morphology of the expanded PBS nanocomposite<br />

foams was observed with an optical microscope (i-Camscope,<br />

Sometech, Korea). The average cell size d was estimated by<br />

measuring the maximum diameter of each cell perpendicular to the<br />

skin from an OM micrograph. To determine the cell size distribution,<br />

the size of at least 75 cells in the OM micrograph was measured. The<br />

cell density Nc and the mean wall thickness δ are given by [36]<br />

N<br />

c<br />

=<br />

1-<br />

( ρ / ρ )<br />

10<br />

f<br />

−4<br />

d<br />

3<br />

p<br />

and ⎟ ⎟<br />

⎛<br />

⎞<br />

⎜ 1<br />

δ = d<br />

− 1<br />

⎜<br />

⎝<br />

1−<br />

( ρf<br />

/ ρ p )<br />

⎠<br />

where ρp, ρf, and d are the pre-foamed density, the post-foamed<br />

density in g/cm 3 , and the average cell size in mm, respectively.<br />

10<br />

)


3. Results and discussion<br />

3. 1. PBS/MWNTs nanocomposites<br />

Figure 1 shows Raman signals of MWNT and nanocomposites<br />

produced by different preparation schemes. Disorder (D) and<br />

tangential (G) bands of MWNT at about 1298 cm -1 and 1582 cm -1 were<br />

distinctly observed at 1064 nm excitation. We confirmed two<br />

characteristic absorption peaks caused by MWNT in the case of the<br />

nanocomposites, indicating that nanocomposites based on MWNTs<br />

were successfully formed regardless of preparation methods.<br />

3. 2. Nanocomposite morphology and dispersion of MWNTs<br />

Figure 2 illustrates the morphology and dispersion of nanocomposites<br />

containing 2 wt% MWNTs, obtained from solution blending, melt<br />

mixing, and the SOAM method, respectively. Overall, MWNTs<br />

randomly dispersed throughout the PBS matrix are shown in Figures<br />

2a, 2b, and 2c. However, in Figures 2a and 2b, bundles of MWNTs are<br />

partially seen, because of their van der Waals interaction, and the pull-<br />

out phenomenon of MWNTs from the PBS matrix also existed due to<br />

the weak interfacial adhesion between the matrix and MWNTs as<br />

11


compared with Figure 2c. Therefore, it can be noted that the<br />

nanocomposites fabricated by the SOAM method showed better<br />

MWNTs dispersion and stronger polymer-MWNTs interfacial force than<br />

those obtained by means of solution blending and melt mixing,<br />

respectively. Figure 2d shows a TEM micrograph of a nanocomposite<br />

with 2 wt% MWNTs produced using the SOAM method. Individual<br />

MWNTs were randomly dispersed within the PBS matrix, and are<br />

oriented toward the ultratoming direction as well. Because they are<br />

affected by a strong shear force during additional melt mixing, shorter<br />

nanotubes can be seen in some locations.<br />

3. 3. Thermal behavior<br />

Figure 3 shows representative TGA thermograms of the PBS/MWNTs<br />

nanocomposites prepared by the SOAM method. It is evident that the<br />

thermal stability was remarkably enhanced with the addition of MWNTs.<br />

This is attributed to not only the uniform dispersion of MWNTs within<br />

the PBS matrix but also the excellent thermal property of the MWNTs<br />

themselves. In case of 5 % weight loss, the degradation temperature of<br />

all nanocomposites produced via the SOAM method exhibited an<br />

12


improvement of approximately 25 o C relative to that of neat PBS (Td =<br />

331.89 o C).<br />

DSC scans were conducted to determine the effect of MWNTs on the<br />

melting and crystallization behavior of PBS nanocomposites. Figure 4<br />

presents the DSC trace of PBS/MWNTs nanocomposites containing<br />

different MWNTs contents, prepared by the SOAM method. In Figure<br />

4a, the neat PBS and its nanocomposites show the double melting<br />

endotherms in a temperature range of 103 ~ 112 o C, respectively.<br />

Some evidence of double and multiple melting behavior has been<br />

previously reported. [37] Especially in the case of PBS and its<br />

composites, these double melt peaks originate from recrystallization.<br />

The first melting point depends on the presence of metastable lamellae,<br />

which have a large surface/volume ratio and fold surfaces with high<br />

surface-free enthalpies during recrystallization. [38] In other words, two<br />

morphologically different crystalline structures existed during the<br />

heating scans. It can be seen that the first melting temperature of the<br />

nanocomposites shifts toward the lower temperature region relative to<br />

that of the neat PBS, indicating that MWNTs play a crucial role in the<br />

reorganization behavior of metastable lamellae. The second melting<br />

temperature, attributed to crystallites with high thermal stability,<br />

13


decreased slightly but increased gradually with increasing MWNT<br />

content. Figure 4b shows the crystallization behavior of the neat PBS<br />

and its composites. The neat PBS was crystallized at 89.55 o C, and<br />

the addition of MWNTs into the PBS matrix led to a shift of the Tc of the<br />

nanocomposites to around 87 o C. This behavior indicates that the<br />

MWNTs do not act as the efficient nucleating agent but impede the<br />

crystallization process of the nanocomposites in all samples.<br />

Consequently, lower temperature was needed to crystallize, as<br />

indicated by the cooling scans. It is shown that the PBS<br />

nanocomposite with 2 wt% MWNTs loading had a higher crystallization<br />

temperature than the other samples.<br />

3. 4. Mechanical testing<br />

The tensile strength and elongation at break curves for the neat PBS<br />

and its composites with different MWNT loadings are shown in Figure<br />

5. The three preparation methods were also compared. It was found<br />

that the nanocomposites prepared by the SOAM method exhibited a<br />

better tensile property than those prepared by solution blending and<br />

melt mixing, respectively. In this study, the ranking of high tensile<br />

mechanical property of the MWNT-reinforced nanocomposites is the<br />

14


SOAM method, the melt mixing method, and the solution blending. In<br />

Figure 5a, it can be observed that the tensile strength of the<br />

nanocomposite increased sharply when a small amount of MWNTs<br />

was added, and gradually achieved a peak value when 2 wt% MWNTs<br />

were loaded, and thereafter decreased with greater MWNT content. It<br />

is believed that the increase in the tensile strength can be ascribed to<br />

the interfacial adhesion between the host polymer matrix and well-<br />

dispersed MWNTs with a high modulus and aspect ratio. In general,<br />

upon the incorporation of the filler into the polymer matrix, the<br />

elongation at break of the nanocomposite was decreased along with<br />

increased tensile strength. However, novel behavior of the elongation<br />

at break of the nanocomposites is shown in Figure 5b. It can seen that<br />

the elongation at break of the nanocomposite was enhanced to a level<br />

higher than that of the neat PBS, and increased up to 2 wt% MWNT<br />

loading, and thereafter decreased with a further increase in MWNT<br />

content. These results could be explained by the MWNTs, having an<br />

outstanding elastic response to deformation, being strongly embedded<br />

within the matrix, thereby inducing efficient load transfer and, in turn, a<br />

homogeneous dispersion.<br />

15


3. 5. Foam behaviors and cell geometry<br />

Figure 7 presents optical micrographs of the cross-sectional surfaces<br />

of the PBS nanocomposite foams with 2 wt% MWNTs content<br />

produced via the SOAM method and foamed for 5 min as a function of<br />

temperature and CBA content. All nanocomposites foamed by CBA<br />

showed a homogeneously closed-cell structure, suggesting that the<br />

presence of well-dispersed MWNTs plays an important role in<br />

increasing melt viscosity of the nanocomposite so as not to destroy the<br />

cell walls, as well as providing the nucleating sites for cell formation.<br />

We note that the cell size increased with more adding the amount of<br />

CBA, also increasing the temperature. If a proper melt viscosity that<br />

restricts the expansion of cells is established, the cell geometry will be<br />

strongly affected by the aforementioned two factors. From the optical<br />

microscopy data, various morphological parameters of PBS/2 wt%<br />

nanocomposite foams obtained by the SOAM method can be<br />

quantitatively determined, as seen in Table 1. The post-foamed mass<br />

density tended to decrease with an increase in both the CBA content<br />

and temperature as a result of expansion of the PBS nanocomposite<br />

foam, while the neat PBS density is 1.268 g/cm 3 . All nanocomposite<br />

foams exhibit a structure ranging from 154.72 μm to 406.30 μm. The<br />

16


cell density, calculated by equation (2), also decreased from 10.67 ×<br />

10 5 cell/cm 3 to 1.09 × 10 5 cell/cm 3 with an increase in both the<br />

blowing temperature and the content of CBA, respectively. These<br />

results are attributed to the well-dispersed MWNTs, which act as a<br />

nucleation agent, facilitating the cell formation process, namely,<br />

heterogeneous nucleation could be achieved at the interfaces between<br />

MWNTs and PBS. Also, the average cell size and cell density were<br />

strongly influenced by the gas evolution as a result of the thermal<br />

decomposition of CBA. The mean cell-wall thickness, based on<br />

equation (3), also decreased in proportion to increased CBA content,<br />

which was attributed to cell growth in accordance with foam expansion<br />

to adjacent cells. We measured the cell size of nanocomposite foams<br />

from optical micrographs and evaluated the distribution of the cell size,<br />

as shown in Figure 8. All foams showed an approximate Gaussian<br />

distribution. Moreover, it can be seen that the lower the blowing<br />

temperature was, the narrower the width of the distribution peaks was.<br />

The narrow peak indicates the high uniformity of cell size, which is<br />

significantly related to the dispersion state of the MWNTs as well as<br />

the applied blowing temperature.<br />

17


Figure 9 illustrates the variation of the blowing ratio in accordance with<br />

blowing temperature, time, and CBA amount. It is seen that the<br />

blowing ratio, which is calculated by equation (1), tends to increase<br />

steadily as the blowing time, temperature and CBA loading are<br />

increased. The maximum values of the blowing ratio at different CBA<br />

loading are ranged from 7.3 to 11.1, indicating that more addition of<br />

CBA in the nanocomposite melt results in the release of more gases to<br />

nanocomposite, and facilitates increased cell size.<br />

18


4. Conclusions<br />

Poly(butylene succinate) (PBS)/MWNTs nanocomposites were<br />

successfully fabricated by three different methods. Via Raman<br />

scattering, we observed two characteristic absorption peaks attributed<br />

to MWNT-OH in the nanocomposites, and verified successful<br />

formation of nanocomposites based on MWNT-OH. The morphology<br />

and dispersion state of the MWNTs in the polymer domain were<br />

examined by SEM and TEM, respectively. The results suggested that<br />

the nanocomposite prepared by the SOAM method exhibited more<br />

dispersed MWNTs and stronger interfacial force with PBS than the<br />

nanocomposites fabricated by means of solution blending and melt<br />

mixing, respectively. The thermal stability was measured by TGA, and<br />

it was found that the thermal stability was enhanced by about 25 o C<br />

relative to neat PBS (Td = 331.89 o C). Double melting endotherms,<br />

which were attributable to a recrystallization process, were observed in<br />

neat PBS and nanocomposites based on MWNTs. In addition, MWNTs<br />

also impeded the crystallization behavior of nanocomposites, as<br />

revealed by DSC measurements. The ranking of high tensile<br />

mechanical property of nanocomposites is the SOAM method, melt<br />

19


mixing, and solution blending. Furthermore, the nanocomposite with 2<br />

wt% MWNTs content showed the highest tensile strength and<br />

elongation at break.<br />

Based on the properties of the nanocomposite described above, we<br />

assumed that the nanocomposite containing 2 wt% MWNTs produced<br />

via the SOAM method would be optimal for the high performance foam.<br />

Well-dispersed MWNTs, which act as a nucleation site for cell<br />

formation, induced an increase in the melt viscosity of the<br />

nanocomposite to restrain cell rupture. As a result, we achieved<br />

nanocomposite foams (>153 μm) with a closed-cell structure by means<br />

of the decomposition of CBA. The morphological parameters such as<br />

the post-foam density, average cell size, cell density, mean cell-wall<br />

thickness, and blowing ratio were significantly affected by the blowing<br />

time, temperature, and CBA content, respectively.<br />

20


5. References<br />

[1] Mohammad Moniruzzaman and Karen I. Winey,<br />

Macromolecules 39, 5194 (2006)<br />

[2] R. Baughman, Science 297, 787 (2002)<br />

[3] X. L. Xie, Y. W. Mai, X. P. Zhou, Mater. Sci. Eng. R 49, 89<br />

(2005)<br />

[4] Jonathan N. Coleman, Umar Khan, and Yurii K. Gun’ko, Adv.<br />

Mater. 18, 689 (2006)<br />

[5] Jason J. Ge, Haoqing Hou, Qing Li, Matthew J. Graham,<br />

Andreas Greiner, Darrell H. Reneker, Frank W. Harris, and<br />

Stephen Z. D. Cheng, J. Am. Chem. Soc. 126, 15754 (2004)<br />

[6] Tianxi Liu, In Yee Phang, Lu Shen, Shue Yin Chow, and Wei-<br />

De Zhang, Macromolecules 37, 7214 (2004)<br />

[7] J. B. Bai, A. Allaoui, Composites: Part A 34, 689 (2003)<br />

[8] Liliane Bokobza, Polymer 48, 4907 (2007)<br />

[9] S. Badaire, P. Poulin, M. Mauggey, C. Zakri, Langmuir 20,<br />

10367 (2004)<br />

[10] Dimitrios Tasis, Nikos Tagmatarchis, Alberto Bianco, Maurizio<br />

Prato, Chemical Reviews 106, 1105 (2006)<br />

21


[11] Kannan Balasubramanian, Marko Burghard, Small 1, 180<br />

(2005)<br />

[12] V. Georgakilas, K. Kordatos, M. Prato, D. M. Guldi, M. Holzinger,<br />

A. Hirsch, J. Am. Chem. Soc 124, 760 (2002)<br />

[13] D. Qian, E. C. Dickey, R. Andrews, T. Rantell, Appl. Phys. Lett.<br />

76, 2868 (2000)<br />

[14] C. Pirlot, I. Willems, A. Fonseca, J.B. Nagy, J. Delhalle, Adv.<br />

Eng. Mater. 4, 109 (2002)<br />

[15] P. Poetschke, A. R. Bhattacharyya, A. Janke, H. Goering,<br />

Compos. Interfaces 10, 389 (2003)<br />

[16] J. Zhu, H. Peng, F. Rodriguez-Macias, J. L. Margrave, V. N.<br />

Khabashesku, A. M. Imam, K. Lozano, E. V. Barrera, Adv. Funct.<br />

Mater 14, 643 (2004)<br />

[17] S. Qin, D. Qin, W. T. Ford, D. E. Resasco, and J. E. Herrera, J.<br />

Am. Chem. Soc 126, 170 (2004)<br />

[18] L. James Lee, Changchun Zeng, Xia Cao, Xiangming Han,<br />

Jiong Shen, Guojun Xu, Composites Science and Technology<br />

65, 2344 (2005)<br />

[19] Suprakas Sinha Ray, Masami Okamoto, Prog. Polym. Sci. 28,<br />

1539 (2003)<br />

22


[20] D. J. Kim, S. W. Kim, H. J. Kang, K.H. Seo, Journal of Applied<br />

Polymer Science 81, 2443 (2001)<br />

[21] Yingwei Di, Salvatore Iannace, Ernesto Di Maio, Luigi Nicolais,<br />

Journal of Polymer Science : Part B : Polymer Physics 43, 689<br />

(2005)<br />

[22] Youhei Fujimoto, Suprakas Sinha Ray, Masami Okamoto,<br />

Akinobu Ogami, Kazunobu Yamada, Kazue Ueda, Macromol.<br />

Rapid Commun. 24, 457 (2003)<br />

[23] L. T. Guan, M. Xiao, Y. Z. Meng, R. K. Y. Li, Polym. Eng. Sci.<br />

46, 153 (2006)<br />

[24] Pham Hoai Nam, Pralay Maiti, Masami Okamoto, Tadao Kotaka,<br />

Takashi Nakayama, Mitsuko Takada, Masahiro Ohshima,<br />

Arimitsu Usuki, Naoki Hasegawa, Hirotaka Okamoto, Polym.<br />

Eng. Sci. 42, 1907 (2002)<br />

[25] M. Hakkarainen, S. Karlsson, A.-C. Albertsson, Polymer 41,<br />

2331 (2000)<br />

[26] Jendrossek, D. Adv In Biochemical Engineering/Biotechnology;<br />

Ghose, T. K.; Fiechter, A., Eds.; Springer: Berlin 71, 293 (2001)<br />

[27] G. Scott, D. Gillead, Degradable Polymers, Chapman & Hall,<br />

London (1995)<br />

23


[28] Lakshmi S. Nair, Cato T. Laurencin, Prog. Polym. Sci. 32, 762<br />

(2007)<br />

[29] Takashi Fujimaki, Polymer Degradation and Stability 59, 209<br />

(1998)<br />

[30] Jian-Hao Zhao, Xiao-Qing Wang, Jun Zeng, Guang Yang,<br />

Feng-Hui Shi, Qing Yan, Journal of Applied Polymer Science 97,<br />

2273 (2005)<br />

[31] Yoshihiro Someya, Toshiyuki Nakazato, Naozumi Teramoto,<br />

Mitsuhiro Shibata, Journal of Applied Polymer Science 91, 1463<br />

(2004)<br />

[32] Sinha Ray S, Yamada K, Okamoto M, Ogami A, Ueda K, Chem<br />

Mater 15, 1456 (2003)<br />

[33] Sinha Ray S, Okamoto K, Okamoto M, Macromolecules 36,<br />

2355 (2003)<br />

[34] Pluta M, Caleski A, Alexandre M, Paul M-A, Dubois P, Journal<br />

of Applied Polymer Science 86, 1497 (2002)<br />

[35] Suprakas Sinha Ray , Mosto Bousmina, Progress in Materials<br />

Science 50, 962 (2005)<br />

[36] Klempner D., Frisch K. C., Handbook of Polymeric Foams and<br />

Foam Technology, Hanser Publishers, Munich, Vienna (1991)<br />

24


[37] Zhaobin Qiu, Motonori Komura, Takayuki Ikehara, Toshio Nishi,<br />

Polymer 44, 7781 (2003)<br />

[38] E. S. Yoo, S. S. Im, Journal of Polymer Science : Part B :<br />

Polymer Physics 37, 1357 (1999)<br />

25


SECTION II. Preparation and Characterization of PBS/Carbon<br />

Nanofiber Nanocomposite Foams<br />

1. Introduction<br />

Polymeric foams are the lightweight materials containing gaseous<br />

voids by means of physical and chemical blowing agents, which have<br />

been mainly used in the insulation, absorbents, cushion, packaging<br />

and weight-bearing structures. [1] For the polymer nanocomposite<br />

foams, the integration of the small amounts of nano-sized fillers into<br />

the virgin polymer may provide enhanced thermal stability and<br />

mechanical properties compared to those of the neat polymer foam,<br />

and can effectively induce the nucleation of a large amount of bubbles<br />

according to the geometry of the added-nanoparticles. [2]<br />

In order to reduce the environmental problems caused by the non-<br />

degradable plastics, the development of the biodegradable polymers<br />

with the properties enough to compare with those of the commodity<br />

plastics greatly have been required and steadily studied by extensive<br />

research groups. Poly(butylene succinate) (PBS), one of the most<br />

applied biodegradable plastics to the scientific and industrial fields, is a<br />

26


thermoplastic, aliphatic polyester produced through polycondensation<br />

reaction of 1, 4-butanediol and succinic acid. It also has melting points<br />

ranging 90 ~ 120 o C, glass transition temperatures ranging -45 ~ -10<br />

o C, and outstanding processability. [3]<br />

Carbon nanofibers (CNFs) are generally known as a carbon allotrope<br />

having diameters of 3 ~ 100 nm with lengths ranging from 0.1 to 1000<br />

µm. Owing to their internal structural characteristics which could be<br />

controlled by the specific synthetic conditions, CNFs display the<br />

chemical stability, high surface area, remarkable electrical and<br />

mechanical properties. So, they have been shown the diverse<br />

applications, for example, the polymer additives for high performance<br />

composites, gas storage materials, electronic components, catalyst<br />

support materials and so on. [4] Especially, in order to act as the<br />

polymer reinforcement, it is necessary to achieve not only the<br />

homogeneous dispersion in the polymer matrix, but the strong<br />

interfacial adhesion between the CNFs and the target polymer. Based<br />

on these issues, the extensive studies related to the CNFs-reinforced<br />

nanocomposites have been carried out. [5-8]<br />

The aim of this study is to prepare the nanocomposite foams with the<br />

closed-cell structures and the high blowing ratio by means of the<br />

27


chemical blowing agent, and to characterize the effect of CNFs acted<br />

as the nucleation sites in the closed-cell system, and the foam<br />

behaviors. Prior to producing the nanocomposite foams, the thermal<br />

and mechanical properties of PBS/CNFs nanocomposites employed by<br />

different fabrication methods were investigated in detail.<br />

28


2. 1. Materials<br />

2. Experimental<br />

PBS (Enpol G4560, MI=1.5 g/10 min), was provided by ire chemical<br />

LTD., Korea. The vapor grown carbon nanofibers called CNF-100 were<br />

obtained from Carbon nano-material technology Co., LTD., Korea.<br />

CNFs have the herringbone structure, and the diameter ranged from<br />

40 to 140 nm and the typical length of 2~10 μm. Chloroform, used as a<br />

solvent in solution blending which was one of the preparation methods<br />

for PBS/CNFs nanocomposite, was supplied by Dongyang chemical,<br />

Korea. The chemical blowing agent (CBA) used was Cellecom ACP-4<br />

grade which was obtained by Kumyang Co., Korea. It consists of a<br />

compound with a specific proportion of azodicarbonamide, N,N'-<br />

dinitrosopenta tetramine, and urea activator.<br />

2. 2. Nanocomposite preparation<br />

To compare the dispersion degree of CNFs in the PBS matrix, we<br />

used three preparation methods such solution blending, melt mixing<br />

and the SOAM method (solution blending and subsequent melt<br />

mixing) [9] , respectively. PBS resin and CNFs were initially dried in<br />

29


vacuo at 60 ºC for 6 h to remove any residual H2O. For solution<br />

blending, PBS pellets were dissolved and stirred in chloroform at 40 ºC<br />

for 5 h using mechanical stirrer. Also, CNFs were dispersed in same<br />

solvent at room temperature for 3 h with spin-bar and then sonicated<br />

for 2 h at 90 W by a ultrasonic bath (Kodo Technical Research Co.,<br />

Japan). After CNFs solution was mixed with PBS solution at 40 ºC for<br />

24 h under vigorous shear condition, nanocomposite casting films were<br />

obtained on rotary hood, and additionally dried at 60 ºC for 24 h to<br />

remove a solvent. For the melt blending, the torque rheometer<br />

(Brabender Plastograph EC, Germany) was used to prepare the PBS<br />

nanocomposite based on a various content of CNFs. The PBS pellets<br />

was melted at 120 ºC for 5 min at a rotary speed of 60 rpm, and then<br />

CNFs were introduced into the mixer, followed by compounding for 10<br />

min. For the SOAM method, the casting films prepared by the solution<br />

blending were firstly cut into the specific size, and then fed into the<br />

torque rheometer and mixed at 120 ºC for 10 min at a rotary speed of<br />

60 rpm. The mass fraction of CNFs ranged from 0.1 to 3 wt%. The<br />

sample codes were named, for example, CNF2S, CNF2M and<br />

CNF2SM, where the numbers represent the mass fraction of CNFs,<br />

30


and the ‘S’ stand for solution blending, ‘M’ for melt mixing, and ‘SM’ for<br />

the SOAM method.<br />

2. 3. Foaming<br />

After the dried solution casting films with different content of CNFs<br />

were fed into the torque rheometer and mixed at 120 ºC for 10 min at a<br />

speed of 60 rpm, ACP-4 varied from 4 to 8 phr was added to the<br />

brabender and subsequently melt-compounded for 10 min. The<br />

samples of the strip type (dimensions 12 mm × 20 mm × 3 mm)<br />

were obtained by using the heating press at 135 ºC for 8 min. Then,<br />

the nanocomposite foams were prepared in forced convection oven at<br />

150, 160 and 170 ºC for 5, 10, 20 and 30 min on the Teflon film.<br />

2. 4. Characterization of nanocomposites and their foams<br />

The thermal gravimetric analysis (TGA) was performed on a TA<br />

Instruments Q50 under air atmosphere using a heating rate of 20<br />

ºC/min from room temperature to 700 ºC. The tensile strength and<br />

elongation at break of nanocomposites were measured by using<br />

Hounsfield Test Equipment at room temperature with the dog-bone<br />

samples. The strain rate was 26 mm/min, and seven measurements<br />

31


were at least performed for each nanocomposite. The dispersion of<br />

CNFs in the PBS matrix was compared by field emission scanning<br />

electron microscopy (FE-SEM) using Hitachi S4300 at 15 kV. The<br />

surface morphology of nanocomposite foams was examined by an<br />

optical microscopy using Sometech i-Camscope. The density of<br />

nanocomposite foams was measured by using Micrometrics Co.<br />

Accupyc 1330 gas pycnometer.<br />

32


3. Results and discussion<br />

Typical TGA thermograms of PBS/CNFs nanocomposites obtained<br />

from different three approaches are given in Figure 1. The introduction<br />

of CNFs with high thermal resistance into the PBS matrix caused the<br />

enhancement of thermal stability in given CNFs content as compared<br />

to the virgin PBS judging from Figure 1(a) and 1(b). The relative<br />

thermal properties of PBS/CNFs nanocomposites were evaluated by<br />

comparing both T95(temperature for 5% weight loss) and<br />

T50(temperature for 50% weight loss) values (Table 1). In general,<br />

PBS/CNFs nanocomposties produced by the SOAM method showed<br />

much higher thermal stability than those produced by solution blending<br />

and melt mixing. Also, PBS nanocomposite with 2 wt% loading of<br />

CNFs prepared by the SOAM method particularly exhibited the highest<br />

T50 value(409 ºC). Therefore, it could be thought that the SOAM<br />

method was the effective way in order to provide a homogenous<br />

dispersion of CNFs in the PBS matrix.<br />

Figure 2 shows the variation of tensile strength and elongation at break<br />

for neat PBS and PBS/CNFs nanocomposites as a function of CNF<br />

loading. As shown in Figure 2(a), tensile strength of the<br />

33


nanocomposites enhanced greatly with the addition of CNFs, and the<br />

CNFs content at maximum tensile strength was only 2 wt% regardless<br />

of the preparation methods. These results are due to the physical<br />

interaction between the PBS matrix and reinforced-CNFs with high<br />

mechanical property, and more uniform dispersion at the 2 wt%<br />

loading of CNFs. However, the tensile strength of nanocomposites<br />

slightly started to decrease after 2 wt%, which attributed to the<br />

aggregation of CNFs in the PBS matrix. It was also found that<br />

nanocomposites prepared by the SOAM method showed relatively<br />

higher tensile strength than those of the other preparation methods.<br />

As the CNFs content increased, the slight decrease in the elongation<br />

at break was observed as indicated in Figure 2(b). This variation was<br />

resulted in part from the modifications in crystalline fraction of the PBS<br />

matrix as more CNFs added. It thus is believed that the CNFs led to a<br />

decrease in ductility and an increase in brittleness.5 In addition, we<br />

can found that the decrease in the elongation at break was less in<br />

nanocomposite obtained from the SOAM method than in those of two<br />

other preparation methods.<br />

The relative state of dispersion of CNFs inside the PBS matrix was<br />

compared by scanning electron microscopy (SEM) as presented in<br />

34


Figure 3, respectively. Figure 3a shows that in nanocomposite<br />

produced by solution blending, individual CNFs were locally<br />

aggregated in the polymer matrix(in circle symbol), and several voids<br />

attributed to a solvent were observed. In the case of nanocomposite<br />

obtained from melt mixing, several CNFs, as shown by arrow in Figure<br />

3b, were pulled out from the PBS matrix during the cryo-fracture,<br />

indicating the poor polymer-CNFs wettability. In contrast, it was found<br />

that not only the majority of CNFs were well-dispersed throughout the<br />

polymer matrix, but also broken rather than being pulled out as<br />

indicated by arrow in Figure 3c. The breakage of CNFs was as a result<br />

of the strong interfacial adhesion between the polymer and CNFs<br />

which acts as a key factor to produce a remarkable mechanical<br />

property. It was also confirmed that the SOAM method plays an<br />

important role to separate the CNFs into individual nanofibers well in<br />

the PBS matrix. Consequently, we believed that the PBS<br />

nanocomposite with 2 wt% CNFs loading fabricated by the SOAM<br />

method was an ideal candidate to make a high performance-<br />

nanocomposite foam by analyzing their thermal and mechanical<br />

properties.<br />

Cross-sectional morphology of nanocomposite foams expanded at 4<br />

35


phr CBA loading was observed via optical microscopy as shown in<br />

Figure 4. All samples showed the oval shape and homogeneous<br />

closed-cell structure. In other words, each cell is surrounded by<br />

connected faces with any ruptures. CBA started to decompose at<br />

temperature higher than 135 ºC and released a gas in the<br />

nanocomposite system, which can cause the bubble nucleation and<br />

growth from the surface of CNFs used as a nucleant. [10,11] In addition,<br />

to achieve the closed-cell system, the nanocomposite should have a<br />

sufficient melt viscosity and strength not so as to break the cell wall. [12]<br />

It thus can be thought that the presence of CNFs served as the<br />

efficient nucleation sites, as well as contributed to produce a closed-<br />

cell structure in foaming process. We also found that the average cell<br />

size was prominently increased as the blowing time and temperature<br />

increased. These two conditions facilitate the decomposition of blowing<br />

agent.<br />

In Table 2, we summarized the morphological parameters of PBS/2<br />

wt% CNFs nanocomposite foaming for 5 min. Average cell size (d) was<br />

calculated by measuring the maximum diameter of each cell<br />

perpendicular to the skin from an OM micrograph, and cell density (Nc)<br />

and mean cell-wall thickness (δ) are given by<br />

36


N<br />

c<br />

=<br />

1-<br />

( ρ / ρ )<br />

10<br />

f<br />

−4<br />

d<br />

p<br />

3<br />

and ⎟ ⎟<br />

⎛<br />

⎞<br />

⎜ 1<br />

δ = d<br />

− 1<br />

⎜<br />

⎝<br />

1−<br />

( ρf<br />

/ ρ p )<br />

⎠<br />

where ρp, ρf, and d are the pre-foamed density, the post-foamed<br />

density in g/cm 3 , and the average cell size in mm, respectively. [13-15]<br />

It was found that the neat PBS density was 1.26 g/cm3, but the post-<br />

foamed density of nanocomposite foam decreased from 0.980 to 0.508<br />

g/cm 3 with increasing the blowing temperature and CBA content. Also,<br />

average cell size increased as the blowing temperature and CBA<br />

loading increased, indicating that the high blowing temperature makes<br />

it easier for the CBA to be decomposed in foaming duration, and the<br />

addition of CBA determines the size and number of bubbles in a<br />

nanocomposite melt. Moreover, we obtained the maximum value of cell<br />

density especially at 6 phr of CBA content at each temperature,<br />

suggesting the existence of a lot of bubbles in the same area. The<br />

mean cell-wall thickness showed a tendency to decrease gradually<br />

according to the further bubble growth.<br />

Figure 5 represents the blowing ratio of PBS/2 wt% CNFs<br />

nanocomposite foams as a function of the blowing temperature, time,<br />

and CBA content. The blowing ratio was calculated as follows: [16]<br />

37


3<br />

the density of neat PBS (1.268 g/cm )<br />

ratio =<br />

the density of PBS/CNFsnanocomposite<br />

foam (g/cm<br />

Blowing 3<br />

We can see that the increase of the blowing temperature, time, and<br />

CBA content led to the blowing ratio increase. The maximum values of<br />

the blowing ratio at 4, 6, and 8 phr CBA content were 5.028, 7.813,<br />

and 1<strong>2.0</strong>25, respectively. It is also believed that the inter-relationship of<br />

above foaming conditions plays a key role to determine the blowing<br />

ratio.<br />

The blowing temperature dependence on the cell size distribution of<br />

PBS/2 wt% nanocomposite foams is represented in Figure 6. We<br />

measured the size of more than 75 cells in fractured surface from the<br />

OM micrograph. It is evident that the CNFs-reinforced nanocomposite<br />

foams almost were in accordance with a Gaussian distribution. When<br />

increasing the blowing temperature, the broad cell size distribution was<br />

examined, indicating that the dispersity of the cell size is much larger.<br />

Also, it was found that the presence of CNFs acted as an efficient<br />

nucleation agent, and well-controlled the cell size accordingly.<br />

38<br />

)


4. Conclusions<br />

Using a chemical blowing agent, PBS/CNFs nanocomposite foams<br />

with a closed-cell structure and high blowing ratio was successfully<br />

obtained. It was found that to produce the high performance-<br />

nanocomposite foams, the optimal CNFs content was 2 wt%, and the<br />

SOAM method was an effective way to disperse the CNFs in the PBS<br />

matrix based on the enhancement of thermal and mechanical<br />

properties for the nanocomposites. In foaming process, when the<br />

decomposition of CBA occurred in a nanocomposite melt, well-<br />

dispersed CNFs acted as the efficient nucleation sites in the PBS<br />

matrix-CNFs interface. Nanocomposite foams exhibited the oval-shape<br />

and the homogeneous closed-cell structures without any ruptures. A<br />

high blowing ratio of about 12 was obtained at 170 ºC for 30 min, as<br />

well as a uniform cell size distribution of nanocomposite foams<br />

observed.<br />

39


5. References<br />

[1] D. Klempner and K. C. Frisch, Handbook of polymeric foams and<br />

foam technology, Hanser Publishers, Munich, Vienna (1991)<br />

[2] L. James Lee, C. Zeng, X. Cao, X. Han, J. Shen, and G. Xu,<br />

Compos. Sci. Technol. 65, 2344 (2005)<br />

[3] T. Fugimaki, Polym. Degrad. Stabil. 59, 209 (1998)<br />

[4] K. P. de Jong and J. W. Geus, Catal. Rev.-Sci. Eng. 42, 481<br />

(2000)<br />

[5] K. Lozano and E. V. Barrera, J. Appl. Polym. Sci. 79, 125 (2001)<br />

[6] K. Lozano, J. Bonilla-Rios, and E. V. Barrera, J. Appl. Polym. Sci.<br />

80, 1162 (2001)<br />

[7] K. Lozano, S. Yang, and Q. Zeng, J. Appl. Polym. Sci. 93, 155<br />

(2004)<br />

[8] Y. K. Choi, K. Sugimoto, S. M. Song, Y. Gotoh, Y. Ohkoshi, and<br />

M. Endo, Carbon 43, 2199 (2005)<br />

[9] S. K. Lim, E. H. Lee, and I. Chin, J. Mater. Res. 23, 1168 (2008)<br />

[10] R. B. McClurg, Chem. Eng. Sci. 59, 5779 (2004)<br />

[11] J. Shen, C. Zeng, and L. James Lee, Polymer 46, 5218 (2005)<br />

[12] S. K. Lim, S. G. Jang, S. I. Lee, K. H. Lee, and I. Chin, Macromol.<br />

40


Res. 16, 218 (2008)<br />

[13] M. Okamoto, P. H. Nam, P. Maiti, T. Kotaka, T. Nakayama, M.<br />

Takada, M. Ohshima, A. Usuki, N. Hasegawa, and H. Okamoto,<br />

Nano Lett. 1, 503 (2001)<br />

[14] P. H. Nam, P. Maiti, M. Okamoto, T. Kotaka, T. Nakayama, M.<br />

Takada, M. Ohshima, A. Usuki, N. Hasegawa, and H. Okamoto,<br />

Polym. Eng. Sci. 42, 1907 (2002)<br />

[15] S. S. Ray and M. Okamoto, Macromol. Mater. Eng. 288, 936<br />

(2003)<br />

[16] L. T. Guan, M. Xiao, and Y. Z. Meng, Polym. Eng. Sci. 46, 153<br />

(2006)<br />

41


SECTION I. Processing of Poly(butylene succinate)(PBS)/Carbon<br />

Nanotube Nanocomposite Foams<br />

Table 1 Morphological parameters of PBS/2 wt% MWNTs<br />

nanocomposite foam obtained by the SOAM method (blowing time = 5<br />

min).<br />

Blowing<br />

temperature/ o C<br />

CBA<br />

content/phr<br />

ρf/g·cm -3 d/μm<br />

42<br />

Nc × 10 -5<br />

/cell·cm -3<br />

δ/μm<br />

4 0.767 154.72 10.67 91.422<br />

150 6 0.690 207.87 5.82 79.955<br />

8 0.616 230.73 4.50 79.582<br />

4 0.607 200.82 5.63 96.622<br />

160 6 0.525 231.61 4.72 70.954<br />

8 0.397 249.75 3.95 69.896<br />

4 0.567 234.50 3.99 95.523<br />

170 6 0.487 320.88 <strong>2.0</strong>8 66.282<br />

8 0.344 406.30 1.09 69.660


SECTION II. Preparation and Characterization of PBS/Carbon<br />

Nanofiber Nanocomposite Foams<br />

Table 1 Thermal stability for PBS/CNFs nanocomposites.<br />

Preparation<br />

method<br />

Solution<br />

blending<br />

Melt<br />

mixing<br />

SOAM<br />

method<br />

CNFs content (wt%) T95 ( o C) T50 ( o C)<br />

0.1 346 408<br />

0.5 325 405<br />

1 325 404<br />

2 327 405<br />

3 349 405<br />

0.1 344 400<br />

0.5 352 404<br />

1 340 400<br />

2 355 406<br />

3 343 403<br />

0.1 350 407<br />

0.5 352 406<br />

1 355 402<br />

2 353 409<br />

3 352 404<br />

43


Table 2 Temperature and CBA content dependence of morphological<br />

parameters.<br />

Blowing<br />

temperature/ o C<br />

CBA<br />

content/phr<br />

ρf/g·cm -3 d/μm<br />

44<br />

Nc × 10 -5<br />

/cell·cm -3<br />

δ/μm<br />

4 0.980 109.30 17.40 120.041<br />

150 6 0.792 119.18 22.18 75.337<br />

8 0.702 148.62 13.60 73.826<br />

4 0.912 126.36 13.91 112.119<br />

160 6 0.748 135.51 16.48 76.099<br />

8 0.716 161.66 10.30 83.356<br />

4 0.860 163.54 7.36 124.764<br />

170 6 0.671 171.97 9.26 78.655<br />

8 0.508 207.23 6.74 60.445


SECTION I. Processing of Poly(butylene succinate)(PBS)/Carbon<br />

Nanotube Nanocomposite Foams<br />

Figure 1 FT-Raman spectra of the PBS nanocomposites with 2 wt%<br />

content prepared by three different methods: (a) MWNT (b) solution<br />

blending (c) melt mixing and (d) SOAM method.<br />

45


Figure 2 SEM images of the PBS nanocomposites with 2 wt% content<br />

prepared by three different methods: (a) solution blending (b) melt<br />

mixing (c) SOAM method, and (d) TEM micrograph of the 2 wt%<br />

PBS/MWNTs nanocomposites prepared by SOAM method.<br />

46


Figure 3 TGA thermograms of the PBS nanocomposites containing<br />

various MWNTs content prepared by the SOAM method.<br />

47


Figure 4 DSC traces of the PBS nanocomposites containing various<br />

MWNTs content prepared by the SOAM method: (a) melting and (b)<br />

crystalline behaviors.<br />

48


Figure 5 Tensile strength and elongation at break curves of the PBS<br />

nanocomposites with various MWNTs content prepared by three<br />

different methods.<br />

49


Figure 6 Optical micrographs of PBS nanocomposite foams with 2<br />

wt% MWNTs content prepared by the SOAM method as a function of<br />

blowing temperature and CBA content. The blowing time was fixed at 5<br />

min.<br />

50


Figure 7 Cell size distribution of PBS nanocomposite foams with 2<br />

wt% MWNTs content prepared by the SOAM method at 150 o C: (a) 4<br />

(b) 6 and (c) 8 phr. The blowing time was fixed at 5 min.<br />

51


Figure 8 Blowing ratio of PBS nanocomposite foams with 2 wt%<br />

MWNTs content prepared by the SOAM method as a function of<br />

blowing time and temperature and CBA content: (a) 4 (b) 6 and (c) 8<br />

phr.<br />

52


SECTION II. Preparation and Characterization of PBS/Carbon<br />

Nanofiber Nanocomposite Foams<br />

Figure 1 TGA thermograms of PBS/CNFs nanocomposites by means<br />

of different preparation methods.<br />

53


Figure 2 The effect of CNFs content and different preparation methods<br />

on the (a) tensile strength and (b) elongation at break.<br />

54


Figure 3 SEM images of PBS/2 wt% CNFs nanocomposite prepared<br />

by (a) solution blending (b) melt mixing and (c) SOAM method.<br />

55


Figure 4 The effect of blowing time and temperature on the cross-<br />

sectional cell morphology of PBS/2 wt% CNFs nanocomposite foams<br />

measured by the optical microscope. CBA content was fixed at 4 phr.<br />

56


Figure 5 Blowing temperature and time dependence on the blowing<br />

ratio of PBS/2 wt% CNFs nanocomposite foams with different CBA<br />

loading: (a) 4 (b) 6 and (c) 8 phr.<br />

57


Figure 6 Cell size distribution of the PBS/2 wt% CNFs nanocomposite<br />

foams as a function of the blowing temperature: (a) 150 (b) 160 and (c)<br />

170 o C. The blowing time and CBA content were fixed for 5 min and at<br />

4 phr, respectively.<br />

58

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!