INTRODUCTION - bictel.ulg.ac.bebictel.ulg.ac.be/.../unrestricted/Naveau_Chapter5.pdf · Chapter 5:...

22
Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites in scCO 2 113 INTRODUCTION The dispersion of layered silicate (clay) in polymer matrices provides materials with enhanced properties (e.g. mechanical, thermal, barrier and fire properties), depending in particular on the extent of individual clay sheets exfoliation at nanoscale in the matrix 1,2 . Among the different methods for the preparation of polymer/clay nanocomposites, in situ polymerization is recognized as a technique of choice to facilitate the exfoliation 3-6 . In this method, the layered silicate is swollen within the liquid monomer or a monomer solution, so that polymer formation can occur inside the clay galleries. It is generally admitted, especially when hydrophobic monomers are concerned, that the in situ strategy requires a pre-treatment of the clay in order to improve its compatibility with the host monomer/polymer and to allow a good final dispersion 6 . This can be achieved by exchanging the inorganic cations of natural clay with suitable organic compounds. These organic cations can act as compatibilizers 7-9 , or they can bear a functional reactive group, such as a monomer 4,10-14 or an initiator 9,11,15-18 . For example, Zhu et al. 13,14 used a home-made ammonium cation containing one styryl group to modify pristine montmorillonite clay and performed in situ radical polymerization of styrene in bulk. They obtained a completely exfoliated nanocomposite with styrene 13 but observed a mixed exfoliated-intercalated nanocomposite with methyl methacrylate (MMA) 14 . Diaconu et al. 11 used a derivative of MMA, 2-methacryloylethyl-hexadecyldimethylammonium as organomodifier, to enhance the compatibility between the clay platelets and an acrylic matrix in waterborne nanocomposites by mini-emulsion polymerization. However, they reported only partial exfoliation, with a preferential location of clay platelets at the surface of polymer particles. In the aim of synthesizing polymers with well-defined architecture and predictable molecular weights, Böttcher et al. 15 used atom transfer radical polymerization (ATRP) with a suitable initiator previously exchanged in the silicate layers for the polymerization of MMA in acetone. This polar solvent enables a good dispersion of organomodified clay but, due to high concentration of monomer and insolubility of polymer, the medium soon becomes viscous which limits the degree of conversion. Recently, supercritical carbon dioxide (scCO 2 ) appeared as an interesting alternative medium for the in situ preparation of polymer/clay nanocomposites 19-25 . Indeed, thanks to the excellent

Transcript of INTRODUCTION - bictel.ulg.ac.bebictel.ulg.ac.be/.../unrestricted/Naveau_Chapter5.pdf · Chapter 5:...

Page 1: INTRODUCTION - bictel.ulg.ac.bebictel.ulg.ac.be/.../unrestricted/Naveau_Chapter5.pdf · Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites

Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites in scCO2

113

INTRODUCTION

The dispersion of layered silicate (clay) in polymer matrices provides materials with enhanced

properties (e.g. mechanical, thermal, barrier and fire properties), depending in particular on

the extent of individual clay sheets exfoliation at nanoscale in the matrix1,2

. Among the

different methods for the preparation of polymer/clay nanocomposites, in situ polymerization

is recognized as a technique of choice to facilitate the exfoliation3-6

. In this method, the

layered silicate is swollen within the liquid monomer or a monomer solution, so that polymer

formation can occur inside the clay galleries.

It is generally admitted, especially when hydrophobic monomers are concerned, that the in

situ strategy requires a pre-treatment of the clay in order to improve its compatibility with the

host monomer/polymer and to allow a good final dispersion6. This can be achieved by

exchanging the inorganic cations of natural clay with suitable organic compounds. These

organic cations can act as compatibilizers7-9

, or they can bear a functional reactive group, such

as a monomer4,10-14

or an initiator9,11,15-18

. For example, Zhu et al.13,14

used a home-made

ammonium cation containing one styryl group to modify pristine montmorillonite clay and

performed in situ radical polymerization of styrene in bulk. They obtained a completely

exfoliated nanocomposite with styrene13

but observed a mixed exfoliated-intercalated

nanocomposite with methyl methacrylate (MMA)14

. Diaconu et al.11

used a derivative of

MMA, 2-methacryloylethyl-hexadecyldimethylammonium as organomodifier, to enhance the

compatibility between the clay platelets and an acrylic matrix in waterborne nanocomposites

by mini-emulsion polymerization. However, they reported only partial exfoliation, with a

preferential location of clay platelets at the surface of polymer particles. In the aim of

synthesizing polymers with well-defined architecture and predictable molecular weights,

Böttcher et al.15

used atom transfer radical polymerization (ATRP) with a suitable initiator

previously exchanged in the silicate layers for the polymerization of MMA in acetone. This

polar solvent enables a good dispersion of organomodified clay but, due to high concentration

of monomer and insolubility of polymer, the medium soon becomes viscous which limits the

degree of conversion.

Recently, supercritical carbon dioxide (scCO2) appeared as an interesting alternative medium

for the in situ preparation of polymer/clay nanocomposites19-25

. Indeed, thanks to the excellent

Page 2: INTRODUCTION - bictel.ulg.ac.bebictel.ulg.ac.be/.../unrestricted/Naveau_Chapter5.pdf · Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites

Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites in scCO2

114

transport properties of scCO2, it is possible to polymerize a monomer inside clay platelets

without any restriction associated with high system viscosity like in bulk or in classical

solvents24

. Besides, this solvent has very attractive characteristics such as low toxicity, non-

flammability, low cost, easy recyclability, accessible critical parameters (31.1°C, 73.8 bar)

and tunable solvent strength26,27

. Moreover, since CO2 is a gas at ambient temperature,

polymer product recovery is straightforward upon depressurization.

The interest in this green solvent for the in situ polymerization in presence of clay started with

the work of Zerda et al.19

. They took advantage of the low viscosity of scCO2 to prepare

PMMA nanocomposites with high ammonium-modified clay loading (20-50 wt%), through

radical polymerization. Subsequently, aliphatic polyesters (poly( -caprolactone)24

and

poly(lactide)28

)/clay masterbatches with a high clay concentration (up to 66 wt%) and a high

degree of intercalation were prepared by Urbanczyk et al. by in situ ring-opening

polymerization of the corresponding lactones. The final nanocomposites are then obtained by

redispersing the masterbatches by melt blending in the polymer matrix of interest. In contrast,

Zhao et al. used low inorganic content for the polymerization of MMA in scCO2. The

particularity of their contribution is the use of fluorinated20

and poly(dimethylsiloxane)-based

ammonium cations21

for the modification of montmorillonite clay. They showed that these

modified clays provide effective steric stabilization and lead to the formation of exfoliated

PMMA nanocomposites.

In the first chapter, we established the use of scCO2 as a medium for clay organomodification.

When pristine clay is contacted with an appropriate organic cation in a high pressure vessel

under supercritical conditions, ionic exchange occurs and after depressurization a ready-to-use

powder is obtained. The process was applied for the preparation of thermally stable

organoclays suitable for melt blending at high temperature (see Chapter 2-4).

In the present chapter, we describe the application of our scCO2 process for clay

organomodification with four different functional organic compounds, i.e. ammonium cations

containing a monomer unit, a fluorinated chain, a polydimethylsiloxane (PDMS)-segment and

an initiator for ATRP. The aim of the study is to examine the potential of those functional

organoclays for the preparation of exfoliated nanocomposites through in situ polymerization

in scCO2, eventually via a one-pot two step convenient process. Methyl methacrylate was

chosen as a model monomer since experimental conditions for its polymerization are already

well established in scCO226,27,29

.

Page 3: INTRODUCTION - bictel.ulg.ac.bebictel.ulg.ac.be/.../unrestricted/Naveau_Chapter5.pdf · Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites

Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites in scCO2

115

EXPERIMENTAL SECTION

1 Materials

Sodium montmorillonite (MMT) labeled as Cloisite® Na

+ (cationic exchange capacity or CEC

of 92.6 meq/100g) was supplied by Southern Clay Products (Rockwood Additives Ltd.).

Methyl methacrylate (MMA, Aldrich, 99 %) was deoxygenated by nitrogen bubbling just

before use. Free radical initiator 2,2′-Azobis(isobutyronitrile) (AIBN, Aldrich, 98 %) and

carbon dioxide (CO2, Air Liquide, 99,998 %) were used as received.

For ATRP experiments, CuBr (Aldrich, 98 %) was purified by dispersion in glacial acetic

acid and stirring for a few hours followed by filtration, washing with acetone and drying

under reduced pressure at 80°C. Fluorinated macroligand (Mn = 12 000 g/mol, 3 TEDETA

units/chain) was synthesized according to a reported method30

.

2 Organomodifiers synthesis

Organic salt 1 (S1, Scheme 1) was obtained by lyophilization of [2-

(methacryloyloxy)ethyl]trimethylammonium chloride solution (Aldrich, 80 wt% in H2O) and

was stored in a dessicator. S2 and S3 were synthesized by quaternization of 2-

(dimethylamino)ethyl methacrylate (Aldrich, 98 %) with respectively 1-bromododecane

(Aldrich, 97 %) and 1H,1H,2H,2H-perfluorooctyl iodide (Aldrich, 96 %) in acetone (50°C

under reflux for 3 days), based on literature description31

. S4 is a polydimethylsiloxane mono-

terminated with a trimethylammonium group (PDMS-N+(CH3)3, Mn 1400 g/mol) described

elsewhere32

and was kindly provided by Céline Labuyère (SMPC, UMons). S5 was obtained

in two steps from 11-bromoundecan-1-ol (Aldrich, 98 %) as previously reported15

.

3 Modification of clay

In a typical example, 1 g of unmodified clay (MMT) and a slight excess of organic cation (1.1

equivalents relative to CEC) were poured in a 40 ml high pressure reactor. 1 ml of ethanol

was added as a co-solvent and the vessel was dipped in an oil bath at 40°C. Following

thermostatization, CO2 was injected with an Isco automatic syringe pump and pressure was

adjusted to 200 bar. After stirring during 2 h, the reactor was depressurized and the modified

clay was dried under vacuum at 50°C overnight to remove the co-solvent. No purification was

performed before polymerization assays. Only to evaluate the yield of exchange, as-obtained

Page 4: INTRODUCTION - bictel.ulg.ac.bebictel.ulg.ac.be/.../unrestricted/Naveau_Chapter5.pdf · Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites

Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites in scCO2

116

organoclays were washed at room temperature firstly with water and secondly with acetone

followed by drying under vacuum at 80°C overnight, as described previously.

4 Polymerization

Free radical polymerizations were conducted in scCO2 in 20 ml high pressure reactors with

4.0 ml of MMA monomer (3.7 g). AIBN (1 wt% compared to monomer) and clay (3 wt% as

inorganics compared to monomer) were first poured in the reactor containing a magnetic

stirring bar. The vessel was dipped in an oil bath at 65°C and oxygen was removed by CO2

venting for 15 min. MMA was then injected under CO2 flow, followed by pressurization at

300 bar.

ATRP experiments were carried out using a similar procedure, with 6.6 ml of MMA (6.2 g),

0.290 g of initiator-modified clay (1.95.10-4

mol of initiator, 3 wt% inorganics to monomer),

0.028 g of CuBr (1.95.10-4

mol) and 0.780 g of fluorinated macroligand (6.50 10-5

mol).

Polymerization reactions proceeded at 300 bar and 65°C for various times (between 4 and 120

h), before cooling in an ice bath and slow release of CO2. Residual monomer was eliminated

by drying under vacuum at 80°C overnight.

5 Characterization

As-obtained organoclays were analyzed by X-ray diffraction (XRD) in reflectance mode with

a powder diffractometer Siemens D5000 (Cu Kα radiation with λ = 0.15406 nm, 50 kV, 40

mA, Ni filter, step size 0.05° and step time 2s) in order to evaluate their interlayer distance.

The organic content before and after washings was determined by thermogravimetric analysis

(TGA, Q500 from TA Instruments) at a heating rate of 20K per min, from room temperature

to 600°C under nitrogen flow.

The morphology of composites was observed by electron microscopy. For transmission

electron microscopy (TEM, Philips CM100 at 100 kV), ultrathin sections (50-80 nm) of

samples were prepared with a Leica EM FC6 ultramicrotome at room temperature and

deposited on copper grids. Powdery samples were previously pressed for 30 s at 180°C or

embedded and cured in an epoxy resin. Samples for scanning electron microscopy (SEM,

JEOL JSM 840-A) were metallized with Pt.

Molecular weights of polymers were determined by gel permeation chromatography (GPC) in

dimethylformamide (DMF) with LiBr (0.025 M) at 55°C (flow rate: 1 ml/min), with a Waters

600 liquid chromatograph equipped with a 410 refractive index detector and styragel HR

Page 5: INTRODUCTION - bictel.ulg.ac.bebictel.ulg.ac.be/.../unrestricted/Naveau_Chapter5.pdf · Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites

Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites in scCO2

117

columns (HR1, 100-5000; HR3, 500-30000; HR4, 5000-500000, HR5, 2000-4000000)

calibrated with poly(methyl methacrylate) standards. Samples were first centrifugated in

DMF, followed by two successive filtrations (through 0.45 nm and 0.20 nm filters). ATRP

samples were extracted from clay sheets by a preliminary exchange with LiCl ions in a

saturated THF solution at room temperature during one night, followed by precipitation in

methanol.

RESULTS AND DISCUSSION

1 Preparation of functional organoclays in scCO2

Organoclays were prepared in scCO2 by ionic exchange between the sodium ions present in

the interlayer space of natural montmorillonite and ammonium cations. For the first time, the

scCO2 process was applied to ammonium cations bearing a vinyl monomer or a

polymerization initiator. Organomodifier structures are presented in Scheme 1. Organic salt 1

(S1, commercially available) consists of a methacrylate moiety and a trimethylammonium

group. It was suggested that, if polymerization could occur with incorporation of the

methacrylate moiety of this clay organomodifier in the growing polymer chains, exfoliation

will be favored4. S2 is similar to S1 except for the substitution of a methyl group by a dodecyl

chain on the nitrogen cation. This modification was indeed proven to be beneficial to expand

clay interlayer distance and hence further facilitate exfoliation33

. S3 and S4 respectively

contain a fluorinated and a PDMS-chain, both known as CO2-philic segments. These

ammonium compounds were tested in reference to the work of Zhao et al.20,21

who

demonstrated that fluorinated or PDMS-clay dispersions are effective stabilizers for

polymerization of MMA in scCO2. Two differences must be noted: our fluorinated segment is

smaller (6 fluorinated carbons compared to 10) and bears a methacrylate end group, while our

PDMS-surfactant is also smaller (15 Si(CH3)2-O units compared to ~44) and is mono-

terminated by a trimethylammonium group to avoid bonding of two adjacent clay sheets.

Lastly, S5 was chosen in reference to the interesting work of Böttcher et al.15

on ATRP. A

few examples exist with this initiator but none in scCO2 to our best knowledge.

All organic salts except for S4 are in the solid state at room temperature and present a poor

solubility in scCO2. Therefore, to favor the exchange reaction, a little amount of ethanol (2.5

vol%) was added as co-solvent (cfr Chapter 1). No co-solvent was added in case of S4. The

Page 6: INTRODUCTION - bictel.ulg.ac.bebictel.ulg.ac.be/.../unrestricted/Naveau_Chapter5.pdf · Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites

Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites in scCO2

118

results are presented in Table 1. For the first two samples, the stoechiometry of ammonium

cation compared to the cationic exchange capacity of clay (CEC) was varied in order to study

its influence on the interlayer distance (d001) and the exchange yield. The latter is determined

by comparing the weight loss due to organic content as measured by TGA of washed products

(classically between 150 and 550°C) to the theoretical weight loss based on CEC and

molecular weight of ammonium cation.

Table 1. Results of organomodification in scCO2

Code Function Stoech.a

Org. cont.b

(wt%)

d001b

(nm)

Yieldc

(%)

MMT-1a monomer

1.1 11.1 1.43 71

MMT-1b 2.2 20.8 1.87 89

MMT-2a monomer +alkyl chain

1.1 24.0 1.70 76

MMT-2b 1.6 30.3 1.74 95

MMT-3 monomer +fluor.chain 1.1 27.2 1.30 35

MMT-4 PDMS 1.1 58.5 2.06 85

MMT-5 ATRP initiator 1.1 31.0 1.80 80

a stoechiometry of ammonium cation compared to the clay cationic exchange capacity (CEC)

bdetermined by TGA on as-obtained dried organoclays

cdetermined on washed, dried organoclays

O

O

N+

C l

O

O

N+

B r

O

O

N+

F

F

F

F

F

F

F

FF

F

FF

F

I

Si

O

Si

O

Si O

N+

15

B r

N+

O

O

B r

B r

Scheme 1. Structure of used organomodifiers (numbers refer to those of Table 1)

S1 S2 S3

S4

S5

Page 7: INTRODUCTION - bictel.ulg.ac.bebictel.ulg.ac.be/.../unrestricted/Naveau_Chapter5.pdf · Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites

Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites in scCO2

119

Figure 1. XRD patterns of functional organoclays prepared in scCO2

Compared to natural clay (interlayer distance d001 of 1.1 nm), all as-obtained organoclays

present an enlarged gallery height, suggesting intercalation of ammonium cations between the

clay layers. MMT-1a (with 10 % excess ammonium) presents a relatively small enlargement

of d001 (from 1.1 to 1.4 nm), which is directly related to the small size of organomodifier 1.

Interestingly, the gallery height may be increased up to 1.9 nm with the addition of twice as

much organomodifier, indicating that excess organic cation can be located between the clay

layers as noticed by other authors34

. A further increase in organic cation concentration did not

lead to further interlayer enlargement. In comparison, S2 with an alkyl chain of 12 carbons

leads to a maximum d001 of 1.7 nm, reached with 1.1 equivalents. A peak shouldering is

nevertheless observed with 1.6 equivalents (MMT-2b), which might suggest at least partial

location of excess ammonium ion in the galleries and the breaking up to some extent of the

organized structure of clay. In both examples, yield of exchange is improved with increasing

organomodifier concentration.

Concerning MMT-3, a distance of 1.3 nm does not seem enough for extensive intercalation of

the corresponding ammonium cation. Indeed, after washing, a large fraction of organic

content is eliminated, leading to a very poor yield. The difficulty to intercalate S3 between

clay layers might be explained by the iodide counterion which is less effective for ionic

exchange in scCO2 compared to bromide and chloride ions. Indeed, we showed in Chapter 1

that the formation of the inorganic salt from the combination of sodium cations of the clay

and anions of organomodifier is the driving force of the ionic exchange in scCO2 and NaI has

the lowest lattice enthalpy, thus less favored formation from its ions under the same

conditions.

Page 8: INTRODUCTION - bictel.ulg.ac.bebictel.ulg.ac.be/.../unrestricted/Naveau_Chapter5.pdf · Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites

Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites in scCO2

120

MMT-4 presents a gallery height of 2.1 nm, which at first might seem a low value relative to

the high molecular weight (1400 g/mol) of the used ammonium salt. In comparison,

Labruyère et al.32

did the ionic exchange in water and obtained a d001 = 1.7 nm, but they did

not exclude that it could correspond to d002, as the peak was not well defined. In fact, if the

peak corresponded to d002, the related d001 would be of 3.4 nm (4.1 nm in our case) which

could still be visible on the spectrum (limit of detection is 2θ = 2° which corresponds to d =

4.4 nm). We rather believe that flexible PDMS-ammonium chains are lying flat on the clay

surface. According to the relatively well-defined peak obtained and the high yield, PDMS-

ammonium chains are thus successfully intercalated in MMT in scCO2 medium, without the

need of a co-solvent. Moreover, a direct proof of successful ionic exchange is detected on

XRD spectra of as-obtained organoclays (Figure 1), by the observation of NaCl or NaBr salt

coming from the recombination of sodium ion from the clay and counter-ion from the

ammonium cation (Cl- or Br

- depending on the ammonium). In contrast, on MMT-3

diffractogram, very thin peaks observed at 2θ=6 and 2θ=23 may be attributed to the non-

exchanged salt.

Finally, initiator-modified clay (MMT-5) was prepared in scCO2. With S5, a much broader

diffraction peak is observed for the as-obtained clay, with central distance of 1.8 nm. The

same organic cation was used by other groups, with different reported gallery heights:

Böttcher et al.15

obtained a relatively sharp diffraction peak with d001 = 1.88 nm while

Behling et al.35

noted a distance of 1.45 nm (no XRD spectra was shown). Larger discrepancy

are even found among the same group: Shipp et al.36,37

once reported a distance of 1.46 nm

and elsewhere 1.96 nm, the latter distance from a weaker and broader diffraction peak

compared to the first. In each case (including our work), the same proportion of

montmorillonite clay to initiator was used. Yields of exchange are claimed between 80 and 88

% but these must be compared carefully as different washing methods were performed.

The observation of diverse X-ray diffraction patterns for organoclays containing the same

organomodifier in identical proportion is linked to the strong θ dependence of experimental

parameters, as well as arrangement and organization of constituents38

. In fact, the differences

in reported interlayer distances may be explained by a late work of Behling et al.39

who

demonstrated a two-step addition of initiator to MMT in acetone (Figure 2). After 48 h, the

interlayer distance measured is 1.45 nm, while 1.85 nm is reached after 96 h. In between (72

h), a broad peak is observed, suggesting a transition between the two intercalation heights.

Page 9: INTRODUCTION - bictel.ulg.ac.bebictel.ulg.ac.be/.../unrestricted/Naveau_Chapter5.pdf · Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites

Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites in scCO2

121

Our obtained pattern corresponds to that “transition state”. We may conclude that ionic

exchange in scCO2 is much faster, as the transition state is reached after 2 h compared to 72 h.

Further investigation is however needed to better understand the phenomenon in wet solvent

(not explained by the authors) and in scCO2.

Figure 2. X-ray patterns of initiator-clay prepared in acetone, showing a two-stage addition of

initiator (S5) to MMT, reproduced with permission from Behling et al.39

This part demonstrated that supercritical carbon dioxide is an effective medium for the

preparation of three types of functional organoclays. Compared to the more classical wet

process, the scCO2 technique allows an easy recovery of organoclays in a simple, fast and

very convenient way. To emphasize this advantage, no washing or purification was carried

out on the organoclays before the polymerization tests.

2 Polymerization in scCO2

Polymerization tests in scCO2 are presented in three sections. First, methacrylate ammonium-

modified clays (MMT-1 and MMT-2) are used for the free radical polymerization of methyl

methacrylate (MMA). Secondly, PDMS-modified clay is introduced for pseudo-dispersion

polymerization and thirdly dispersion atom transfer radical polymerization from initiator-clay

is presented. In each paragraph, the morphology of nanocomposites is studied.

2.1 Free radical polymerization with methacrylate-bearing clays

Clays modified by methacrylate-bearing ammonium (MMT-1b and MMT-2b) were used in

the free radical polymerization of MMA in scCO2. The quality of clay dispersion in the

polymer matrix was then evaluated, in order to establish organomodifier structure/degree of

exfoliation relationships.

Page 10: INTRODUCTION - bictel.ulg.ac.bebictel.ulg.ac.be/.../unrestricted/Naveau_Chapter5.pdf · Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites

Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites in scCO2

122

Free-radical polymerization in scCO2 was carried out at 65°C for 16 h at 300 bar with

magnetic stirring. Quantity of monomer was 4 ml in a 20 ml reactor, with 1 wt% AIBN and 3

wt% inorganics (both with respect to monomer).

Table 2. Results of free radical polymerization of MMA in the presence of organomodified

clays in scCO2

clay time

(h)

conversion

(%)

Mn

(g/mol)

Mw

(g/mol) Mw/Mn

sample

morphology

1 MMT-1b 16 80 not determined flakes

2 MMT-2b 16 83 46400 257000 5.5 flakes

Conditions: 65°C, 300 bar, 700 rpm, 1 wt% AIBN, 3 wt% inorganics

The visual observation of samples 1 and 2 reveals an aggregated yellowish powder at the

bottom of the reactor covered by white solid flakes to the top. A TGA analysis (not shown)

indicates that the aggregated powder contains a large majority of inorganics. We must thus

conclude that clay was not mixed with MMA during polymerization. This is also confirmed

by XRD analysis (not shown) where no enlargement of initial organoclay interlayer distance

is observed. A possible explanation is the fast precipitation of growing PMMA chains in

scCO2, forming a heterogeneous system and leaving no time for the soluble monomer to

penetrate inside the clay layers.

To help clay swelling with MMA and facilitate further polymerization with incorporation of

the methacrylate moiety of this clay organomodifier in the growing polymer chains, we added

a previous step of “clay soaking” at a temperature below the decomposition temperature of

AIBN, namely 40°C, and a pressure of 180 bar. After 2 h, the temperature was increased to

65°C, which consequently increased the pressure to 350 bar. The samples were collected after

7 h under the polymerization conditions and the conversion obtained was around 40 % with

both organoclays. Again, visual observation of these samples shows a large heterogeneity,

with a phase-separation between clay and PMMA. Clay agglomerates (yellowish) are eye-

detected on pressed disks (Figure 3).

Figure 3. Pressed disks of PMMA synthesized in scCO2, left: without clay, middle: with

MMT-1b, right: with MMT-2b

Page 11: INTRODUCTION - bictel.ulg.ac.bebictel.ulg.ac.be/.../unrestricted/Naveau_Chapter5.pdf · Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites

Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites in scCO2

123

An intercalated morphology was reported in the literature for polymerization in bulk of MMA

with clay modified by cations bearing a vinyl group4. However, to our best knowledge, no

example exists in scCO2 with monomer-modified clay and further investigation is needed to

better understand the phenomenon.

2.2 Free radical polymerization with PDMS-clay

Polydimethylsiloxane (PDMS)-modified clay can find application in the polymerization of

MMA in scCO2 because of the CO2-philic character of the siloxane chains. If the insoluble

clay platelets get attached to the growing PMMA chains, the soluble PDMS-organomodifier

can help providing steric stabilization of PMMA in this medium. This technique was referred

to as pseudo-dispersion polymerization by Zhao et al.21

.

Polymerization in scCO2 was performed at conditions close to those described by Zhao, i.e. at

65°C and 300 bar for 4 h with magnetic stirring. Quantity of monomer was 4 ml in a 20 ml

reactor, with 1 wt% AIBN and 3 wt% inorganics (both with respect to monomer).

Table 3. Results of free radical polymerization of MMA with PDMS-clay in scCO2

sample clay time

(h)

conversion

(%)

Mn

(g/mol)

Mw

(g/mol) Mw/Mn

sample

morphology

4 MMT-4 4 30 52400 150000 2.9 powder

Conditions: 65°C, 300 bar, 700 rpm, 1 wt% AIBN, 3 wt% inorganics

After 4 h of reaction, 30 % of conversion is reached and a homogeneous fine white powder is

collected from the reactor, suggesting a stabilizing effect of PDMS-clay (containing PDMS-

monotelechelic trimethyl ammonium, cfr Scheme 1). This effect is consistent with the

observation of Zhao et al. who modified clay with an - bis(aminopropyl)-terminated

PDMS quaternized with HCl21

. The PDMS content to monomer in our study and in the work

of Zhao were calculated and compared in Table 4. An equal amount of PDMS to monomer

was used.

Table 4. PDMS content in PMMA nanocomposites compared to that of Zhao et al.21

wt% of PDMS-clay

to monomer

wt% of PDMS

in PDMS-clay

wt% of PDMS

to monomer

this study 7.2* 58.5 4.2

Zhao21

6.0** 65.0 3.9

*corresponding to inorganic content of 3 wt%, **lowest % to obtain a fine powder

Page 12: INTRODUCTION - bictel.ulg.ac.bebictel.ulg.ac.be/.../unrestricted/Naveau_Chapter5.pdf · Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites

Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites in scCO2

124

SEM analysis reveals that the PMMA/PDMS-clay nanocomposites primarily consist of

spherical PMMA particles, with an average particle size around 10 µm (Figure 4), same as

that obtained by Zhao.

Figure 4. SEM picture of PMMA with PDMS-clay (Table 3)

The proposed mechanism of stabilization is firstly based on description by DeSimone et al.27

for dispersion polymerization in scCO2. They state that steric stabilization of a colloidal

dispersion is usually imparted by amphiphilic macromolecules that become adsorbed onto the

surface of the dispersed phase. These amphiphilic macromolecules contain an anchoring

segment, which attaches to the particle usually by physical adsorption, and stabilizing

moieties that are soluble in the continuous phase. The stabilizing moieties project in the

continuous phase and prevent flocculation by mutual excluded volume repulsion. Secondly,

Lagaly et al.40

have shown that clay platelets can be used as stabilizers for emulsions and

inverse emulsions by encapsulation of droplets through formation of a three-dimensional

inorganic network. Thus, combining these two statements, it can be suggested that soluble

PDMS chains, linked to the clay platelets, may act as the soluble stabilizing moieties, while

the clay platelet itself acts as the anchoring segment, being attached to the C=O group of

methyl methacrylate by hydrogen bonding with remaining interlayer water molecules (Figure

5). Hydrogen bond interactions between PMMA and clay were indeed detected previously by

FT-IR20

.

Figure 5. Schematic illustration of stabilization of PMMA particle (in blue) by clay platelets

(in green) bearing PDMS chains (in orange), adapted from21

Page 13: INTRODUCTION - bictel.ulg.ac.bebictel.ulg.ac.be/.../unrestricted/Naveau_Chapter5.pdf · Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites

Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites in scCO2

125

The morphology of the powdery nanocomposites was then investigated by X-ray diffraction

and transmission electron microscopy. XRD pattern of PMMA containing PDMS-clay

(MMT-4) is a smooth line with no visible peak below 2θ = 10 (Figure 6). The disappearance

of the characteristic peak of the organoclay might suggest an exfoliation in the matrix.

However, as commonly admitted, this must be confirmed by other techniques, such as TEM.

Figure 6. XRD pattern of PDMS-clay and PMMA nanocomposite with PDMS-clay

TEM analysis of pressed sample reveals an extensive destructuration of the clay stacks, with

preferential arrangement of clay platelets (individual and small stacks) in ribbons, as shown

on Figure 7. This observation confirms the proposed mechanism of stabilization, with the

location of clay on the surface of the PMMA particles.

Zhao et al. reported a uniform dispersion of silicate layers in the PMMA matrix, however this

statement was based on TEM only and the distribution of the silicate layers depends on the

angle of observation.

.

Figure 7. TEM pictures of PMMA with PDMS-clay from pressed sample (30 s at 180°C)

500 nm

Page 14: INTRODUCTION - bictel.ulg.ac.bebictel.ulg.ac.be/.../unrestricted/Naveau_Chapter5.pdf · Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites

Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites in scCO2

126

As a conclusion on free radical polymerization in presence of clay in scCO2, we observed that

MMA-like clays were not homogeneously dispersed in PMMA, probably due to a fast

precipitation of growing polymer chains. On the contrary, PDMS-like clay was successfully

exfoliated in PMMA, thanks to a stabilizing effect of siloxane chains, leading to pseudo-

dispersion polymerization with formation of spherical PMMA microparticles stabilized by

PDMS modified clay located at their surface. This process is particularly interesting as no

supplementary surfactant is needed. Moreover, as the organoclay itself is prepared in scCO2,

the preparation of stabilized PMMA nanocomposites could be processed in one-pot in future

applications, after optimization of the degree of conversion. The analysis of thermal,

mechanical and fire properties of these nanocomposites still needs to be carry out. Finally, it

is noteworthy that PDMS-clay may also find applications in the preparation of silicon

polymer nanocomposites.

2.3 Atom transfer radical polymerization with initiator-modified clay

Another interesting route to obtain fully dispersed polymer/layered silicate nanocomposites is

the use of initiator-modified clay. Indeed, the anchoring of an initiator inside the clay layers is

the best way to ensure polymer chain growth from the silicate with a progressive delamination

during reaction. We chose the ATRP-initiator developed by Böttcher et al.15

because of its

simple synthesis in mild conditions. Moreover, atom transfer radical polymerization is a very

convenient process to synthesize polymers with well-defined architecture and molecular

weights. Böttcher et al. obtained very interesting results regarding control of MMA

polymerization in the presence of initiator-modified clay. However, they worked in acetone

with a limited monomer conversion ( 50 %). Detailed morphology of samples was also not

provided. Our objective was to test the same initiator for ATRP of MMA in scCO2, for its

advantageous low viscosity and high mass transport capacity.

DeSimone et al.41

studied the effect of using different ligands in the copper-mediated ATRP

in scCO2. They obtained highest yields and best control when a fluorinated ligand was used

and attributed this to the enhanced catalyst solubility in scCO2. This is why CO2-philic

fluorinated ligands with tetraethyldiethylenetriamine (TEDETA) as complexing group were

synthesized and used in this study, i.e. a fluoroalkyl-substituted ligand (f-L) and a fluorinated

macromolecular ligand (f-ML) sketched in Scheme 2. The latter was chosen because of its

demonstrated dual role, (i) the complexation of the ATRP catalyst (copper salt) and (ii) the

stabilization of the growing polymer particles (PMMA)29,42

. At the end of the polymerization,

Page 15: INTRODUCTION - bictel.ulg.ac.bebictel.ulg.ac.be/.../unrestricted/Naveau_Chapter5.pdf · Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites

Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites in scCO2

127

PMMA microspheres can be recovered in powder form30,42

. For further use, the macroligand

may either be eliminated through supercritical fluid extraction or serve to produce a material

with hydrophobic properties.

NN

N

O

Et

Et Et

Et

O

C8F

17

[ ( )n ( )m ]p

NN

N

O

O

Et

Et Et

Et

O

OOO

C8F

17

HO2C

S H

[ ( )n ( )m ]p

[ ( )n ( )m ]p

NN

N

O

O

Et

Et Et

Et

O

OOO

C8F

17

HO2C

S H

[ ( )n ( )m ]p

Scheme 2. General scheme of fluorinated ligand (f-L, left) and fluorinated macroligand (f-

ML, right, 12 000 g/mol, 3TEDETA units/chain) used for ATRP

ATRP of MMA was initiated in scCO2 by an ammonium-terminated bromoisobutyrate (cfr

Scheme 1), alone or anchored to clay layers by electrostatic interaction. The weight

percentage of clay to monomer was calculated to be of 3 % as inorganics (i.e. MMT-5 as-

obtained organoclay with 31 wt% of initiator). An equivalent proportion of initiator, catalyst

(CuBr) and ligand was used. Conditions were a temperature of 65°C and a pressure of 300

bar. Quantity of monomer was 6.6 ml in a 20 ml reactor. After 70 h of magnetic stirring at

700 rpm, the experiment carried out without clay (entry a, Table 5) reached a good conversion

while the ones containing clay (entries b and c, Table 5) went to around 50 % conversion.

These were then repeated for longer times (5 days, entries d and e, Table 5). Assays d and e

are done in the same conditions to test reproducibility.

Table 5. Results of ATRP polymerization in scCO2

clay ligand time

(h)

conv.

(%)

Mnth

(g/mol)

MnGPC

(g/mol) Mw/Mn f

sample

morphology

a - f-ML 70 94 29900 45000 1.75 0.66 fine powder

b MMT-5 f-L 70 53 16900 35200 1.54 0.48 flake

c MMT-5 f-ML 70 54 17200 39900 2.85 0.43 flake

d MMT-5 f-ML 120 97 30800 146000 1.63 0.21 flake + powder

e MMT-5 f-ML 120 95 30200 89800 1.69 0.34 fine powder

Conditions: 65°C, 300 bar, 700 rpm, [initiator]/[TEDETA] = 1, [CuBr]/[TEDETA] = 1, Mnth

= [MMA]0/[initiator]0 x MwMMA x conv., f = Mnth/MnGPC

Page 16: INTRODUCTION - bictel.ulg.ac.bebictel.ulg.ac.be/.../unrestricted/Naveau_Chapter5.pdf · Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites

Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites in scCO2

128

A first general observation is the slower polymerization rate in scCO2 compared to

polymerization in solution (conversion of 50 % reached after 4 h in 2 ml acetone with 3 ml

MMA and same initiator-clay, Böttcher et al.15

), mainly due to the higher dilution used in

supercritical medium. Secondly, the lower conversion observed with the presence of clay may

be explained by the necessity for the catalyst, complexed by the ligand, to enter the gallery

space to meet the initiator. Yet, with its fluorinated chains, the ligand has more affinity for

CO2 then for silicate. As a result, the initiation step might be delayed.

The relatively large polydispersity without clay is probably linked with the poor solubility of

the initiator bearing an ammonium-end group in scCO2, inducing inhomogeneous initiation of

reaction. As a reference, polydispersity obtained by Böttcher et al.15

in acetone is 1.1.

Comparing assay a to assays d and e, the polydispersity is not influenced by the presence of

clay. It is however higher with the macroligand compared to that with the smaller ligand,

most likely due to its size which hinders the access to initiator molecules localized inside the

clay galleries. The suggested mechanism is a start of polymerization through the externally

located initiator molecules with a progressive delamination of clay. As the clay gallery height

increases, more internally located initiator molecules can be reached.

Concerning the initiator efficiency (f = Mnth/MnGPC), it is limited to 0.65 without clay, which

is consistent with previous reports for ATRP of MMA in scCO2, with methyl α-

bromophenylacetate initiator30

. When anchored to silicate layers, the efficiency is decreased.

The result calculated with fluorinated ligand (0.47) is close to the value calculated from the

results of Böttcher et al. (0.52) with the same ammonium-initiator but in acetone. This could

be attributed to the non accessibility of part of initiator molecules located inside clay galleries,

due to insufficient clay sheet delamination. An interesting study published recently reported

the influence of graft density on kinetics of surface-initiated ATRP of polystyrene from

montmorillonite39

. With the same initiator as in present study but in bulk, the efficiency is of

95 %, whatever the graft density. Following above hypothesis of insufficient delamination,

their high efficiency can be explained, at least partially, by the use of ultrasonication to

enhance MMT stacks dispersion. Besides, they eliminated termination effects by the

introduction of excess CuBr2, as was done previously for ATRP from silicon surface43

. The

use of ultrasonication and CuBr2 should thus be considered to optimize our system in scCO2.

Page 17: INTRODUCTION - bictel.ulg.ac.bebictel.ulg.ac.be/.../unrestricted/Naveau_Chapter5.pdf · Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites

Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites in scCO2

129

Another interesting perspective of these preliminary tests would be to combine the small

ligand, to facilitate penetration between the clay layers, and the macroligand to ensure further

stabilization of the system. Indeed, stabilization of PMMA particles was effective with the use

of the macroligand, as proved by the collection of a fine powder with free initiator without

clay (Table 5, entry a) and silicate-anchored initiator (Table 5, entry e). Careful attention

should however be paid to the stirring effectiveness, since insufficient stirring might

explained the lack of reproducibility between samples c, d and e. These samples were

analyzed by scanning electron microscopy (SEM).

Figure 8. SEM pictures of ATRP samples with macroligand (letters refer to entry in Table 5;

a- left: without clay, e-right: with clay)

SEM analysis of sample without clay (Figure 8, sample a) reveals typical spherical particles

with an average size of 5 µm, coexisting with larger spherical particles up to 20 µm,

consistently with previous results obtained for MMA dispersion polymerization in scCO2 with

the same stabilizer and methyl α-bromophenylacetate as ATRP initiator30,42

. In the presence

of clay (Figure 8, sample e), spherical particles of large size distribution coexist with particles

of various forms.

XRD analysis was carried out on sample powder to evaluate the quality of clay dispersion. No

peak is detected in PMMA/clay spectrum (Figure 9), which may be an indication of

exfoliation. However, we should be careful about this statement since the clay itself did not

show a well-defined peak. Further analysis is thus required to prove that a nanocomposite is

obtained. In this optic, ultrathin sections of as-obtained samples were observed by TEM

(Figure 10).

e a

Page 18: INTRODUCTION - bictel.ulg.ac.bebictel.ulg.ac.be/.../unrestricted/Naveau_Chapter5.pdf · Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites

Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites in scCO2

130

Figure 9. XRD pattern of initiator-clay and PMMA nanocomposite with initiator-clay

Figure 10. TEM pictures of PMMA with ATRP-initiator clay (sample e, in epoxy resin)

The TEM images of PMMA powder embedded in epoxy resin reveal the presence of

aggregates and stacks of medium and small size. Individual platelets were also observed,

together with large zones free of clay, indicating only partial intercalation and small extent of

delamination. It is consistent with the difficult access of the macroligand to the initiator inside

the clay platelets. We may assume that when all monomer is consumed, clay aggregates

subsist with intercalated initiator molecules that did not take part to the polymerization

process.

Finally, the morphology of PMMA particles obtained by ATRP may be compared with that of

PMMA particles obtained with PDMS-clay. In both cases, spherical particles of micrometric

size are obtained, with at least partial exfoliation of clay layers. In fact, the two methods show

distinct advantages: the use of a CO2-philic organomodifier (PDMS-ammonium) enables the

clay itself to stabilize the PMMA particles. This method is very simple and easy to implement

but the possibility to control the size and architecture of PMMA is quite limited. On the

5000 nm 500 nm

Page 19: INTRODUCTION - bictel.ulg.ac.bebictel.ulg.ac.be/.../unrestricted/Naveau_Chapter5.pdf · Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites

Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites in scCO2

131

contrary, ATRP will offer the possibility to control the molecular weights but therefore it will

require a careful choice of initiator and ligand, as well as adjustment of other processing

parameters (e.g. stirring, temperature, pressure, quantity of catalyst). Once the ATRP product

being optimized, the thermal and mechanical properties of PMMA nanocomposites should be

analyzed and compared with those of PMMA nanocomposites obtained by free radical

polymerization, to provide supporting information on the interest of each method.

CONCLUSIONS

Three types of functional organoclays were successfully prepared in scCO2. Intercalation of

organomodifier in montmorillonite clay was demonstrated by X-ray diffraction and

polymerization of MMA in presence of these organoclays in scCO2 is reported. While

methacrylate-bearing clays were not homogeneously dispersed in PMMA synthesized by free

radical polymerization, exfoliated PDMS-clay platelets served as effective stabilizer for the

synthesis of this polymer in scCO2, leading to the formation of spherical microparticles with

clay located at their surface. Thirdly, ATRP-initiator modified clay in presence of catalyst and

fluorinated macroligand also led to the formation of PMMA microspheres in scCO2, thanks to

the stabilizing properties of the macroligand. The process is convenient since the polymer is

recovered as a fine powder after synthesis and further optimization of the ATRP process

would allow preparing polymers with well-defined molecular weight. Moreover, functional

cations of various structures should be tested to enhance clay exfoliation in this or another

matrix. Finally, as both clay organomodification and polymerization occur in scCO2, a one-

pot process can be easily envisaged.

Page 20: INTRODUCTION - bictel.ulg.ac.bebictel.ulg.ac.be/.../unrestricted/Naveau_Chapter5.pdf · Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites

Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites in scCO2

132

REFERENCES

1. Alexandre, M.; Dubois, P. Mater. Sci. Eng., R 2000, 28, 1-63.

2. Ray, S. S.; Okamoto, S. Progr. Polym. Sci. 2003, 28, 1539-1641.

3. Weimer, M. W.; Chen, H.; Giannelis, E. P.; Sogah, D. Y. J. Am. Chem. Soc. 1999,

121, 1615-1616.

4. Su, S.; Wilkie, C. A. J. Polym. Sci.,: Part A: Polym. Chem. 2003, 41, 1124-1135.

5. Di, J.; Sogah, D. Y. Macromolecules 2006, 39, 1020-1028.

6. Naveau, E.; Detrembleur, C.; Jérôme, C.; Alexandre, M. Rec. Pat. Mater. Sci. 2009, 2,

43-49.

7. Gilman, J. W.; Awad, W. H.; Davis, R. D.; Shields, J.; Harris, J. R. H.; Davis, C.;

Morgan, A. B.; Sutto, T. E.; Callahan, J. H.; Trulove, P. C.; DeLong, H. C. Chem.

Mater 2002, 14, 3776-3785.

8. Fornes, T. D.; Hunter, D. L.; Paul, D. R. Macromolecules 2004, 37, 1793-1798.

9. Wheeler, P. A.; Wang, J.; Mathias, L. J. Chem. Mater. 2006, 18, 3937-3945.

10. Akelah, A.; Moet, A. J. Mater. Sci. 1996, 31, 3589-3596.

11. Diaconu, G.; Micusík, M.; Bonnefond, A.; Paulis, M.; Leiza, J. R. Macromolecules

2009, 42, 3316-3325.

12. Zeng, C.; Lee, L. J. Macromolecules 2001, 34, 4098-4103.

13. Zhu, J.; Morgan, A. B.; Lamelas, F. J.; Wilkie, C. A. Chem. Mater. 2001, 13, 3774-

3780.

14. Zhu, J.; Start, P.; Mauritz, K. A.; Wilkie, C. A. Polym. Degrad. Stab. 2002, 77, 253-

258.

15. Böttcher, H.; Hallensleben, M. L.; Nuss, S.; Wurm, H.; Bauer, J.; Behren, P. J. Mater.

Chem. 2002, 12, 1351-1354.

16. Fan, X.; Xia, C.; Fulghum, T.; Park, M. K.; Locklin, J.; Advincula, R. C. Langmuir

2003, 19, 916-923.

17. Mittal, V. J. Coll. Interface Sci. 2007, 314, 141-151.

18. Velten, U.; Shelden, R. A.; Caseri, W. R.; Suter, U. W. Macromolecules 1999, 32,

3590-3597.

19. Zerda, A. S.; Caskey, T. C.; Lesser, A. J. Macromolecules 2003, 36, 1603-1608.

20. Zhao, Q.; Samulski, E. T. Macromolecules 2005, 38, 7967-7971.

21. Zhao, Q.; Samulski, E. T. Polymer 2006, 47, 663-671.

22. Yan, C.; Ma, L.; Yang, J. J. Appl. Polym. Sci. 2005, 98, 22-28.

23. Li, J.; Xu, Q.; Peng, Q.; Pang, M.; He, S.; Zhu, C. J. Appl. Polym. Sci. 2006, 100, 671-

676.

24. Urbanczyk, L.; Calberg, C.; Stassin, F.; Alexandre, M.; Jérôme, R.; Jérôme, C.;

Detrembleur, C. Polymer 2008, 49, 3979-3986.

Page 21: INTRODUCTION - bictel.ulg.ac.bebictel.ulg.ac.be/.../unrestricted/Naveau_Chapter5.pdf · Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites

Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites in scCO2

133

25. Sun, F.; Pan, Y.; Wang, J.; Wang, Z.; Hu, C.; Dong, Q. J. Macromol. Sci., Part A

2009, 46, 37-45.

26. Kendall, J. L.; Canelas, D. A.; Young, J. L.; DeSimone, J. M. Chem. Rev. 1999, 99,

543-564.

27. DeSimone, J. M.; Maury, E. E.; Menceloglu, Y. Z.; McClain, J. B.; Romack, T. J.;

Combes, J. R. Science 1994, 265, 356-359.

28. Urbanczyk, L.; Ngoundjo, F.; Alexandre, M.; Jérôme, C.; Detrembleur, C.; Calberg,

C. Eur. Polym. J. 2009, 45, 643-648.

29. Grignard, B.; Jérôme, C.; Calberg, C.; Jérôme, R.; Detrembleur, C. Eur. Polym. J.

2008, 44, 861-871.

30. Grignard, B.; Jérôme, C.; Calberg, C.; Jérôme, R.; Wang, W.; Howdle, S. M.;

Detrembleur, C. Chem. Commun. 2008, 314-316.

31. Zheng, X.; Jiang, D. D.; Wang, D.; Wilkie, C. A. Polym. Degrad. Stab. 2006, 91, 289-

297.

32. Labruyère, C.; Monteverde, F.; Alexandre, M.; Dubois, P. J. Nanosci. Nanotechnol.

2009, 9, 2731-2738.

33. Fornes, T. D.; Yoon, P. J.; Hunter, D. L.; Keskkula, H.; Paul, D. R. Polymer 2002, 43,

5915-5933.

34. Xie, W.; Gao, Z.; Pan, W. P.; Hunter, D.; Singh, A.; Vaia, R. Chem. Mater. 2001, 13,

2979-2990.

35. Behling, R. E.; Cochran, E. W. Polym. Prepr. 2007, 48, 427-428.

36. Argoti, S. D.; Reeder, S.; Zhao, H.; Shipp, D. A. Polym. Prepr. 2002, 43, 267-268.

37. Zhao, H.; Argoti, S. D.; Farrell, B. P.; Shipp, D. A. J. Polym. Sci., Part A: Polym.

Chem. 2004, 42, 916-924.

38. Vaia, R. A.; Liu, W. J. Polym. Sci., Part B: Polym. Phys. 2002, 40, 1590-1600.

39. Behling, R. E.; Williams, B. A.; Staade, B. L.; Wolf, L. M.; Cochran, E. W.

Macromolecules 2009, 42, 1867-1872.

40. Lagaly, G.; Reese, M.; Abend, S. Appl. Clay Sci. 1999, 14, 83-103.

41. Xia, J.; Johnson, T.; Gaynor, S. G.; Matyjaszewski, K.; DeSimone, J. Macromolecules

1999, 32, 4802-4805.

42. Grignard, B.; Jérôme, C.; Calberg, C.; Jérôme, R.; Wang, W.; Howdle, S. M.;

Detrembleur, C. Macromolecules 2008, 41, 8575-8583.

43. Matyjaszewski, K.; Miller, P. J.; Shukla, N.; Immaraporn, B.; Gelman, A.; Luokala, B.

B.; Siclovan, T. M.; Kickelbick, G.; Vallant, T.; Hoffmann, H.; Pakula, T.

Macromolecules 1999, 32, 8716-8724.

Page 22: INTRODUCTION - bictel.ulg.ac.bebictel.ulg.ac.be/.../unrestricted/Naveau_Chapter5.pdf · Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites

Chapter 5: Functional organoclays and routes for the in situ preparation of polymer nanocomposites in scCO2

134