Vous êtes sur la page 1sur 7

Home Search Collections Journals About Contact us My IOPscience

Controlled synthesis of magnesium hydroxide nanoparticles with different morphological

structures and related properties in flame retardant ethylene–vinyl acetate blends

This content has been downloaded from IOPscience. Please scroll down to see the full text.

2004 Nanotechnology 15 1576

(http://iopscience.iop.org/0957-4484/15/11/035)

View the table of contents for this issue, or go to the journal homepage for more

Download details:

IP Address: 155.97.178.73
This content was downloaded on 20/06/2014 at 16:21

Please note that terms and conditions apply.


INSTITUTE OF PHYSICS PUBLISHING NANOTECHNOLOGY
Nanotechnology 15 (2004) 1576–1581 PII: S0957-4484(04)78923-0

Controlled synthesis of magnesium


hydroxide nanoparticles with different
morphological structures and related
properties in flame retardant
ethylene–vinyl acetate blends
Jianping Lv, Longzhen Qiu and Baojun Qu1
State Key Laboratory of Fire Science, Department of Polymer Science and Engineering,
University of Science and Technology of China, Hefei, Anhui 230026,
People’s Republic of China

E-mail: qubj@ustc.edu.cn

Received 2 April 2004, in final form 23 August 2004


Published 1 October 2004
Online at stacks.iop.org/Nano/15/1576
doi:10.1088/0957-4484/15/11/035

Abstract
Magnesium hydroxide nanoparticles with different morphological structures
of needle-, lamellar- and rod-like nanocrystals have been synthesized by
solution precipitation reactions of alkaline with magnesium chloride in the
presence of complex dispersants and characterized in terms of morphology,
particle size, crystal habits and thermal behaviour by transmission electron
microscopy, x-ray diffraction and thermogravimetric analysis. The sizes and
morphologies of magnesium hydroxide nanocrystals can be controlled
mainly by the reaction conditions of temperature, alkaline-injection rate and
the concentrations of reactants. The data show that the needle-like
morphology is of size 10 × 100 nm2 , the lamellar shape 50 nm in diameter
and estimated 10 nm in thickness, and the rod-like nanoparticles 4 µm in
length and 95 nm in diameter, respectively. All three kinds of nanoparticles
are of hexagonal structures. The needle- and lamellar-like nanoparticles can
be obtained by the reactions of alkaline injected into magnesium chloride
solution at about 2 and 20 ◦ C, respectively, while the rod-like nanoparticles
can be prepared by a slower alkaline-injection rate and lower aqueous
ammonia concentration at about 10 ◦ C. The results obtained from the
ethylene–vinyl acetate nanocomposites blended with the lamellar-like
nanoparticles show that magnesium hydroxide nanocrystals possess higher
flame retardant efficiency and mechanical reinforcing effect by comparison
with common micrometre grade magnesium hydroxide particles.

1. Introduction applications in electronics, catalysis, ceramics, nanostructured


composites or halogen-free flame retardants [1]. For
The synthesis of magnesium hydroxide (MH) nanoparticles example, MH nanorods could be used as precursors for the
with special morphologies has recently attracted much synthesis of MgO nanorods [2], which are expected to have
attention because of the expectation of novel properties and novel mechanical, catalytic and electronic properties due
to their extremely small size, large anisotropy and perfect
1 Author to whom any correspondence should be addressed.
crystallinity [3]. MH nanoneedles and nanolamellas could be

0957-4484/04/111576+06$30.00 © 2004 IOP Publishing Ltd Printed in the UK 1576


Controlled synthesis of magnesium hydroxide nanoparticles in ethylene–vinyl acetate blends

good candidates for functional polymeric composites and fibre sulfate in the weight ratio of 1:1 were put into a 250 ml three-
hybrid materials as reinforcing agents or halogen-free flame- necked flask. Secondly, 20 g 25 wt% ammonia water solution
retardants [4, 5]. However, the synthesis of MH nanoparticles was injected into the above flask for 1 h, and then 58 g 8 wt%
with special morphologic structure is difficult, because sodium hydroxide aqueous solution for 2 h, under vigorous
nanoparticles may adopt various shapes [6]. Some research stirring at 2 ◦ C. The mixture was continuously stirred for
effort has been devoted to the synthesis of MH nanoparticles 1 h, and then heated to 40 ◦ C for 2 h before being cooled to
with special morphology. Recently, Li and Ding, and their co- room temperature. The resultant suspension was filtered and
workers, proposed a route to synthesize MH nanorods [7, 8] washed with water to remove the residual impurities. The final
based on the hydrothermal reaction of pure magnesium powder product was dried for 24 h at 80 ◦ C, yielding a white needle-like
in an autoclave at higher temperature using a long reaction MH powder (7.2 g, 98% yield based on magnesium chloride
time and low reactant concentration. Similarly, Yu and hexahydrate).
co-workers produced MH nanoplates by the hydrothermal
The preparation of lamellar-like MH nanoparticles is
reaction of magnesium oxide crystals [9]. Henrist and co-
similar to that of the needle-like MH nanoparticles except
workers produced nanoplates by a wet precipitation method
for using higher reaction temperature of 20 ◦ C. Rod-like MH
using dilute aqueous solution [10]. All these chemical
nanoparticles were obtained at 10 ◦ C under the condition of a
syntheses were, however, characterized by problems such
very slow alkaline-injection rate, first 100 g 5 wt% ammonia
as being limited to one or two morphological preparations,
water solution for 3 h, then 58 g 8 wt% sodium hydroxide
relatively low yields and being economically unacceptable. A
simple fabrication method of MH nanoparticles with several aqueous solution for 4 h.
morphological structures is still a challenge. Polymeric nanocomposites were obtained by the melt
In this paper, we report a solution precipitation method blending method. EVA copolymer (containing 28% vinyl
to prepare well-dispersed MH nanocrystals with different acetate, Sumitomo Chemical Co., Ltd) was mixed with MH
morphologies by using gelatin and lauryl sodium sulfate as nanoparticles for 15 min at 130 ◦ C using a Rheomixer XSS-300
complex water-soluble polymer dispersants. The important in 60 g-batches. After mixing, the samples were hot-pressured
features of this method are that it is relatively simple, effective into sheets of 3 mm thickness under 10 MPa for 3 min at 150 ◦ C
and size-controllable. Three morphological structures with for the limiting oxygen index (LOI) tests, or 2 mm thickness
needle-, lamellar- and rod-like nanocrystals can be obtained for the tensile strength (TS) measurements.
by controlling the reaction temperature, alkaline-injected rate X-ray diffraction (XRD) analysis was carried out using
and reactant concentration, respectively. The flame-retardant a Philips X’ Pert PRO SUPER apparatus (Nicolet Instrument
and mechanical properties of the ethylene–vinyl acetate (EVA) Co., USA) with Cu Kα (λ = 1.5418 Å) radiation at a scan rate
copolymer/MH nanocomposites are also investigated in the of 0.0167◦ s−1 . A transmission electron microscope (TEM,
present study. HITACHI H-800) was used to observe the morphology of
MH nanocrystals. The dried MH product was added into
2. Experiment distilled water to form a suspension, which stabilized by a small
amount (<0.2 wt%) of polyacrylamide (MW = 100 000)
MH nanocrystalline samples were synthesized by a solution for preventing the nanoparticles from agglomerating. The
chemical process so-called homogeneous precipitation in the specific surface area was measured according to the Brunauer–
presence of complex water-soluble polymer dispersants. The Emmett–Teller (BET) method, employing nitrogen adsorption
reaction formulae are given as follows: at liquid nitrogen temperature, by means of a Micromeritics
ASAP 2000 apparatus (Micromeritics Co., USA). MH samples
MgCl2 ·6H2 O + 2NH3 → Mg(OH)2 + 2NH4 Cl + 6H2 O (approximately 0.2 g) were first outgassed for about 2 h at
150 ◦ C under a residual pressure lower than 91.0 Pa, then
MgCl2 ·6H2 O + 2NaOH → Mg(OH)2 + 2NaCl + 6H2 O.
for about 1.5 h at 250 ◦ C under a residual pressure lower
than 2.6 Pa. Thermogravimetric analysis (TGA) was carried
In this method, MgCl2 ·6H2 O was used as a magnesium
out in air using a STA409C thermogravimetric analyser. A
precursor. NH3 and NaOH aqueous solutions were used
10 ± 0.5 mg sample was examined under an air flow rate
as precipitators. The dispersants were used for controlling
of 6 × 10−5 m3 min−1 at temperatures ranging from room
the shape and size by preventing the nanocrystals from
temperature to 800 ◦ C. FTIR spectra were recorded using a
agglomerating. The pH value was monitored during the
precipitation reaction. The alkaline solution was injected Nicolet MAGNA-IR 750 spectrometer.
into the magnesium chloride solution by a peristaltic pump LOI is defined as the volume fraction of oxygen in an
at different speeds. The use of a peristaltic pump provided oxygen–nitrogen atmosphere that will support steady candle-
a better way to control the nucleation step, and to investigate like burning of a material. It has been widely applied as
systematically the effects of reagent concentration, reaction an indicator of polymer flammability. The LOI value was
temperature and drop-wise speed on the morphology and size measured using a ZRY-type instrument (made in China) on
range of MH nanoparticles. sheets 120 × 60 × 3 mm 3 according to the ASTM D2863-77
The fabrication of needle-like MH nanocrystals can be standard. The TSs were measured with a Universal Testing
described as follows. First, a 50 g 50 wt% magnesium chloride Machine (DCS 5000, SHIMADZU, Japan) at 25 ± 2 ◦ C. The
hexahydrate aqueous solution and a 8 g 10 wt% complex crosshead speed was 25 mm min−1 and the dumb-bell shaped
dispersants aqueous solution of gelatin and lauryl sodium specimens were prepared according to ASTM D412-87.

1577
J Lv et al

Table 1. MH nanocrystallines obtained under different reaction conditions.


NH3 8 wt% NaOH
injected injected Specific
Initial NH3 Sizes surface
Sample Reactant concentration Rate Time Rate Time Temp. (nm) area
code injected (wt%) (g h−1 ) (h) (g h−1 ) (h) (◦ C) Morphology by TEM (m2 g−1 )
1 Alkali 25 20 1 29 2 2 Needle-like 10 × 100 58.2
2 Alkali 25 20 1 29 2 20 Lamellar 10 × 50 48.8
3 Alkali 5 33 3 14 4 10 Rod-like 95 × 4000 41.7

3. Results and discussion

It has been proved that the addition of a strong base into a metal
salt solution made it difficult to generate monodispersed metal
hydrous oxides [11]. In the synthesis of MH nanoparticles,
aqueous ammonia was initially used as a weak base to
control the morphology and size of the MH crystal, whereas
the 8 wt% sodium hydroxide aqueous solution as a strong
base was subsequently added to raise the yield of MH.
The morphologies, specific surface areas and sizes of MH
nanocrystals obtained under different reaction conditions are
given in table 1.
It can be seen from table 1 that the temperature has
a key important effect on the formed morphologies of MH
nanoparticles. The MH crystals obtained at the reaction
temperature 2 ◦ C (sample 1) took needle-like morphology in
the size of 10 × 100 nm 2 , as shown in figure 1(a). When the
reaction temperature was raised to 20 ◦ C, the obtained MH
crystals (sample 2) show more regular lamellar shape with
50 nm diameter and estimated 10 nm thickness, as shown
in figure 1(c). This can be interpreted as being due to the
equal speed of crystal growth in plane at higher temperature.
The BET information showed that the obtained MH particles
with needle-like morphology (sample 1) in lower reaction
temperature had a bigger surface area of 58.2 m2 g−1 than
that of 48.8 m2 g−1 with plate-like morphology (sample 2) in
higher reaction temperature.
It should be pointed out that the additional ageing of
the final product at higher temperature had little influence
on the nanocrystals in our experiments. When the reaction
temperature was as low as 2 ◦ C, the resulting nanoparticles
(sample 1) still present good crystallization, as shown in
figure 1(a). There was no change found from the corresponding
selected-area electron diffraction (SAED) patterns after ageing Figure 1. TEM images of MH nanocrystals: (a) and (b), needle-like
for 12 h at 80 ◦ C, which means that the subsequent morphology, synthesized at 2 ◦ C and 3 h alkaline-injection rate with
ageing process at higher temperature could not improve the or without complex dispersants. (c) and (d), lamellar-like
crystallization. A possible explanation is that the very slow morphology, synthesized at 20 ◦ C and 3 h alkaline-injection rate
with or without complex dispersants. (e) and (f), rod-like
injection speed of alkaline solution (at least 3 h for sample 1 morphology, synthesized at 10 ◦ C, lower initial NH3 concentration
and 2, or 7 h for sample 3) in the presence of the complex and 7 h alkaline-injection rate with or without complex dispersants.
dispersants can avoid the rapid aggregation of MH nuclei and
thus result in good crystallization even at the low temperature
of 2 ◦ C. volumes of the longer gelatin chain or by stabilizing the
It is very interesting that the lower initial aqueous growing particle with surfactants or dispersants under the
ammonia concentration and longer injection time of alkaline condition of longer injection time of alkaline solution.
solution such as 7 h at 10 ◦ C can result in the formation of Although preparing nanoparticles with rod structure is not so
rod-like nanoparticles 4 µm in length and 95 nm in diameter novel today, especially for materials with layered structure, the
(sample 3 in table 1). The SAED pattern indicates that appearance of a nanorod structure in such a simple solution
MH nanorods have more perfect crystallization as shown precipitation reaction is still exciting.
in figure 1(e). Presumably, the rod-like particle growth Surfactants or water-soluble polymeric dispersants have
was controlled by restricting particle formation to confined already been used as protecting agents in the synthesis of

1578
Controlled synthesis of magnesium hydroxide nanoparticles in ethylene–vinyl acetate blends

2854 (101)
3698
2916
(001)

(110)
(102)
Transmittance

intensity(a.u.)
a (100) (111) (103) (201)

4000 3500 3000 2500 2000 1500 1000 500


-1
Wavenumbers (cm ) 10 20 30 40 50 60 70 80
2 theta (degree)
Figure 2. FTIR spectrum for sample 2 with lamellar-like
morphology coated with the complex dispersants. Figure 3. XRD patterns of MH nanocrystals with different
morphologies: (a) needle-like, I001 /I110 = 1.7, (b) lamellar-like,
I001 /I110 = 2.3, (c) rod-like, I001 /I110 = 2.4.
nanoparticles with different morphologies [12–14]. In the
present study, we find that complex dispersants play an
important role in the synthesis of three MH morphological Table 2. Crystal data and crystalline size information from the
structures. Figure 1 also compares the TEM images of three XRD patterns of MH.
morphological MH nanoparticles obtained in the presence and Miller Full width at Crystal
absence of the complex dispersants, respectively. It can be Sample Peak position index half-maximum size
seen that the nanocrystals obtained in the absence of complex code (d value nm) (hkl) (deg) (nm)
dispersants tend to agglomerate completely into a large mass 1 0.4765 010 1.04 7.7
of particles, as shown in figures 1(b), (d) and (f). This is 0.2362 101 0.73 11.4
because the very small nanocrystals possess very large surface 0.1568 110 0.42 21.4
energy, which leads the nanocrystals to aggregate in order to 2 0.4765 010 0.64 12.4
lower their surface energy during crystal growth in the absence 0.2362 101 0.55 15.1
of the complex dispersants. This agglomeration is, however, 0.1568 110 0.41 22.0
prevented by the complex dispersants coating the nanocrystals, 3 0.4765 010 0.58 13.7
as shown in figures 1(a), (c) and (e). 0.2362 101 0.50 16.6
0.1568 110 0.40 22.5
Evidence from the surfactants adsorbed on the surface
of MH nanoparticles can be obtained by the FTIR spectrum.
Figure 2 presents the FTIR spectrum of sample 2 in a KBr
pellet. The peaks at 2916 and 2854 cm −1 have been assigned XRD peaks by means of Scherrer formula [15]. The calculated
to the symmetric vibrations of aliphatic groups (–CH2 –)n , crystal sizes are listed in table 2. It is noted that the crystallite
which indicates that the surfactants with the aliphatic groups size, based on the full width at half-maximum of different
are adsorbed on the surface of the MH nanoparticles. The diffraction peaks, have different values. For example, 7.7 nm
sharp and strong peak at 3698 cm−1 is due to the O– (001), 11.4 nm (001) and 21.4 nm (001) for sample 1, which
H stretching vibrations in the MH crystal structures. The indicates that the particles have a morphology of thin plate or
overlapped adsorption peaks between 1440 and 1650 cm−1 needle shape with layers or lengths in the [001] direction. This
can be attributed to the O–H stretching mode in water and the is supported by the TEM studies in figure 1(a). It can be seen
CH2 /NH scissoring mode in the surfactants. The strong peak from figure 3(c) that the rod-like morphology exhibits a higher
at about 440 cm−1 is assigned the Mg–O stretching vibrations I001 /I110 ratio of 2.4 than that of 2.3 and 1.7 for lamellar- and
in MH. needle-like morphologies, respectively, which means a more
Figure 3 shows the XRD patterns (2θ scan) of three pronounced orientation of the rodlets towards the incident x-
morphological structures of MH nanoparticles. All the peaks ray radiation. The preferential orientation is impossible in the
can be indexed to hexagonal MH crystals with lattice constants isotropic case of ‘sand rose’ morphology with the I001 /I110
comparable to the values of JCPDS 7-239. The SAED patterns ratio of 0.51, as reported in the literature [10].
in figures 1(a), (c) and (e) also give positive evidence that The TGA-DTA profiles of sample 2 with lamellar
the nanocrystals with different morphologies are of hexagonal shape are shown in figure 4. It can be seen that one
structures. pronounced weight loss step appears in the temperature
There are no parasite peaks in the XRD patterns of range of 250–396 ◦ C with a corresponding endothermic
figure 3, which indicates that MH nanoparticles obtained by peak observed near 354 ◦ C, which can be ascribed to the
the solution precipitation method in this study are relatively decomposition of MH. The observed value of MH weight
pure. Moreover, the peaks of the three samples are significantly loss is 30.1%, which is nearly equal to the theoretical value
broadened. This indicates that the MH particles have very 30.8%. The two additional weight loss steps near 79.1 and
small grain sizes, which can be estimated from the broadened 216.7 ◦ C are also observed from figure 4, which are due to the

1579
J Lv et al

130
o
0.000 Table 3. Comparison of MH on mechanical and flammable
354.05 C
properties of three kinds of EVA/MH blends.
120
-0.002
Nanosized Micrometre grade Micrometre grade
110 MH MH particles MH particles A MH particles B
Weight Loss (%)

10 × 50 nm2 1–2 µm 2–5 µm

Diriv.Weight
-0.68% -1.40% -0.004 filler
100
level
90 -0.006 (phr) TS (MPa) LOI (%) TS (MPa) LOI (%) TS (MPa) LOI (%)
-30.05%
80
80 10.4 26 9.7 25 9.4 25
-0.008
100 10.6 38 8.6 31 8.2 30
70 120 11.0 42 7.2 36 7.1 37
-0.010 150 17.0 47 7.1 41 6.9 41
60
0 100 200 300 400 500 600 700 800
o
Temperature ( C)
and from 9.4 to 6.9 MPa for micrometre grade MH B (2–
Figure 4. TGA and DTGA curves for sample 2 with lamellar-like
5 µm) composites as additive levels increase from 80 to
morphology in air at a heating rate of 10 ◦ C min−1 .
150 phr, respectively. In contrast, the corresponding values
for nanoparticles composites increase step by step from 10.4
release of absorbed water on the surface of MH and pyrolysis of to 17.0 MPa. This means that the nanosized MH particles
coated organic dispersants, respectively. Their weight losses of possess apparently enhanced mechanical properties.
absorbed water and coated dispersants are determined as being It was noticed that the tensile strength is quite high for the
0.68% and 1.4%, respectively. Therefore, the coated materials nanosized MH particles at high MH filler level, while the LOI is
on the MH nanoparticles can be evaluated as 1.4 wt%. not affected that significantly in table 3. A precise mechanistic
In recent years, increasing attention has been paid interpretation is beyond the scope of this paper, but it can be
to the use of halogen-free flame-retardants in the wire postulated that the process of endothermic decomposition of
and cable industry because the halogen-containing materials MH plays a crucial role in the flame-retardant mechanism.
can cause some problems, such as toxicity, corrosion and An endothermic decomposition of hydrated mineral filler
smoke. Common micrometre grade MH has been used as such as MH reduces the temperature, releases an inert gas
halogen-, smoke- and toxic-free flame retardant additives in the to dilute the combustion gas, and thus prevents the polymer
EVA blends for many years. However, its fatal disadvantages from burning. As the size of MH particles decreases rapidly
are low efficiency as a flame retardant additive and very from micrometre grade to nanosized grade, the character of
high additive level, which causes the mechanical properties endothermic decomposition may not be affected greatly, as
of polymeric materials to decrease sharply. Nanometre described in figure 4. It seems that the particle size or shape
grade MH could possibly solve these problems because of does not appear to be too important to the flame retardancy
the mechanical and flame-retardant reinforcing functions of from figure 4. However, nanosized MH particles with plate-
nanosized MH in polymeric composite materials. Table 3 like morphology are preferred in the circumstance of high filler
lists the LOI and TS data obtained from the three kinds of level, which may increase their flame retardancy by forming
EVA/MH composites. The data show that the LOI values a compact oxide layer, and possess enhanced mechanical
of nanosized MH samples are much higher than those of the property simultaneously.
common micrometre grade MH A (1–2 µm) and B (2–5 µm)
samples in all additive levels, especially for the high additive 4. Conclusions
levels. When the additive levels reach above 100 phr, the LOI
values of the nanosized MH composites are 6–7 higher than The MH nanoparticles with needle-, lamellar- and rod-like
those of the corresponding micrometre-sized MH composites. morphological structures have been prepared successfully by
The above data indicate that nanoparticles preserve higher a solution precipitation method in the presence of complex
flame-retardant efficiency than that of micrometre grade MH. dispersants. Reaction temperature, reactant concentration
The enhancement of the flame-retardant property is due to and alkaline injecting rate are crucial factors in determining
the good dispersion of MH nanoparticles in the EVA matrix the morphologies of the MH nanoparticles. The data from
and the formation of a compact oxide layer in the process XRD and SAED show that all the MH nanocrystals are
of endothermic decomposition of MH nanoparticles with of hexagonal structures. The sizes and morphologies of
lamellar-like morphological structure. This oxide layer, which magnesium hydroxide nanocrystals can be controlled mainly
is similar to the structure of ceramics, may insulate the polymer by the reaction conditions of temperature, alkaline-injected
and trap and oxidize soot precursors, and thus prevents the rate and concentrations of reactants. The lamellar-like
materials from burning [16]. nanocrystals with 50 nm in diameter and 10 nm in thickness
In the context of mineral filled composites, mechanical can be obtained at 20 ◦ C and other suitable conditions, while
properties are inevitably concerned besides their flammable the needle-like nanocrystals with the size of 10 × 100 nm3
characteristics. Many research results showed that the high obtained 2 ◦ C. The rod-like nanocrystals of 95 × 4000 nm3
weight fraction of MH filler additive level in polymers can can be obtained at 10 ◦ C by using the lower initial aqueous
adversely affect some mechanical properties [17–19], and ammonia concentration of 5 wt% and prolonging the injecting
often decrease their tensile strength [20]. It can be seen from time of alkaline solution to 7 h.
table 3 that the values of TS decrease steadily from 9.7 to The TEM images show that complex dispersants play an
7.1 MPa for micrometre grade MH A (1–2 µm) composites, important role in the controlled synthesis of MH nanoparticles

1580
Controlled synthesis of magnesium hydroxide nanoparticles in ethylene–vinyl acetate blends

and the nanocrystals without the complex dispersants result in [4] Xie R C and Qu B J 2001 Preparation of needle-shaped
the agglomeration. The FTIR spectrum gives the evidence of Mg(OH)2 nanoparticles Patent Application Number
00135436 of P. R. China
complex dispersants, while the TG analysis gives its amount
[5] Alexandre M, Beye G, Henrist C, Cloots R, Rulmont A,
of 1.4 wt% on the MH nanoparticles. Jérôme R and Dubois Ph 2001 Macromol. Rapid Commun.
The data from the LOI and mechanical measurements 22 643
demonstrate that the MH nanoparticles in the EVA/MH [6] Spahr M E, Bitterli P, Nesper R, Müller M, Krumeich F and
nanocomposites show better flame-retardant efficiency and Nissen H U 1998 Angew. Chem. 110 1339
[7] Ultamapanya S, Klabunde K J and Schlup J R 1991 Chem.
reinforced mechanical property than common micrometre
Mater. 3 175
grade MH. [8] Li Y D, Sui M, Ding Y, Zhang G, Zhuang J and Wang C 2000
Adv. Mater. 12 818
Acknowledgments [9] Yu J C, Xu A W, Zhang L Z, Song R Q and Wu L 2004
J. Phys. Chem. B 108 64
[10] Henrist C, Mathieu J P, Vogels C, Rulmont A and Cloots R
The authors gratefully acknowledge the support of the National 2003 J. Cryst. Growth 249 321
Natural Science Foundation of China, No. 50373039, and [11] Matijević E 2001 Chem. Mater. 13 412
the China National Key Basic Research Special Funds project [12] Zhan J, Yang X, Wang D, Li S, Xie Y and Qian Y T 2001 Adv.
2001 CB409600. Mater. 12 1348
[13] Burke N A D, Stŏver H D H and Dawson F P 2000 Chem.
Mater. 14 4752
References [14] Sun Y and Xie Y 2002 Adv. Mater. 14 833
[15] Cuttity B D 1978 Elements of X-ray Diffraction 2nd edn
[1] Xie Y, Huang J, Li B, Liu B and Qian Y T 2000 Adv. Mater. 12 (Reading, MA: Addison-Wesley) p 555
1523 [16] Qiu L Z, Xie R C, Ding P and Qu B J 2003 Compos. Struct. 62
[2] Mckelvy M, Sharma R, Chizmeshya A G, Carpenter R W and 391
Streib K 2001 Chem. Mater. 13 921 [17] Wang Z Z, Qu B J, Fan W C and Huang P 2002 J. Appl. Polym.
[3] Lieber C M, Morales A M, Sheehan P E, Wong E W and Sci. 81 206
Yang P 1996 Chemistry on the nanometer scale Proc. [18] Ulutan S and Gilbert M 2000 J. Mater. Sci. 35 2115
Robert A 1996 Welch Foundation 40th Conf. on Chemical [19] Li Z Z and Qu B J 2003 Polym. Degrad. Stab. 81 401
Research (Houston, TX) [20] Nielsen L E 1967 J. Compos. Mater. 1 100

1581

Vous aimerez peut-être aussi