Вы находитесь на странице: 1из 36

Chapter 9

EVA/Clay Nanocomposites Transport Features

Summary
228 Chapter 9

9.1. Introduction

Polymer-clay nanocomposites are hybrid composite materials consisting of

a polymer matrix with dispersed clay nano particles. Nano clays have been

widely used as an inorganic reinforcement for polymer matrices with nano

scale dispersion of the inorganic phase within the polymer matrix [1-4]. The

typical feature size of each filler platelet is approximately 1nm in thickness,

and 100-500nm in length. Mechanical properties are improved due to the

reinforcing effect of the particles [5-7], whereas the thermal stability is

increased [8-9] and the thermal expansion coefficient is reduced [10]. The

large surface to volume ratio of the nano fillers suggests that the particles

may affect the segmental mobility of the polymer phases provided that the

polymer molecules are efficiently attracted to the filler particles in a way

similar to that of amorphous chain segments to crystals in a semi

crystalline polymer.

Polymer-clay nanocomposites have attracted the attention of many

researchers [11-13]. Drozdav et al. [14] showed that the diffusion process

becomes anomalous with higher clay content. Hedengvist et al. [15] studied

the diffusion of methanol through spray dried cheese whey protein-

montmorillonite nanocomposites. Musto et al. [16] examined the diffusion of

water and ammonia through polyimide-silica nanocomposites. Merkel et al. [17]

reported an increase in permeability by adding nano structural fumed silica

to several glassy high free volume polymers. Valsaveld et al. [18]

prepared polyamide- silicate nanocomposites and analysed the influence


EVA/Clay Nanocomposites Transport Features 229

of silicate concentration on the diffusion characteristics and the mechanical

properties. Shantaii et al. [19] prepared polyimide-based nanocomposites

and their properties were characterised by kinetics of water uptake.

Aminabhavi and co-workers [20] used poly (vinyl alcohol)-iron oxide

nanocomposite membranes for pervaporation. High permeability, good

selectivity and stability are the important factors in choosing suitable

pervaporation membranes.

The substantial decrease in moisture permeability was reported for nano

clay-polyamide composites by Okada et al. [21]. The permeability

performance of nano composite normally depends on the clay content,

aspect ratio and the degree of dispersion of silicate layers [22]. The

reduction in vapour permeability was attributed to the extremely high

aspect ratio of clay platelets, which increased the tortuosity of the path of

gas or vapour molecules as it diffuses into the nanocomposite. The wide

application of membranes for gas separation has attracted many polymer

technologists to synthesis new polymeric membranes with good

permeability and selectivity [23]. Polymer layered nanocomposites have

attracted many scientists due to the dramatic improvements in the gas

barrier properties of polymers [24, 25].

In this study EVA/clay nanocomposites containing different filler loading

have been prepared. The nano clay used was closite Na+ which has no

organic modifier. Transport of aromatic hydrocarbons, pervaporation of

chloroform-acetone mixtures, permeation of chlorinated hydrocarbon


230 Chapter 9

vapours and gases like O2 and N2 were investigated. Morphology of

nanocomposites were analysed by XRD and TEM. Positron annihilation

lifetime spectroscopic analysis (PALS) has been used to estimate the free

volume of nanocomposites.

9.2. Results and Discussion


A) Transport of aromatic hydrocarbons through EVA/clay
nanocomposites
9.2.1. Morphology
9.2.1.1. X-ray diffraction analysis (XRD)

XRD is widely used for the characterization of the structure of layered

silicate and polymer nanocomposites. The change in the d-spacing of the

polymer nanocomposite is observed from the position of the peaks in the

XRD patterns in accordance with the well known Braggs equation.

n = 2d Sin (9.1)

where n is an integer which gives the order of reflection, is the wave

length, d is the d-spacing and is the angle of diffraction. X-ray diffraction

method has been used to characterise the formation and structure of

polymer-silicate hybrids by monitoring the position, shape and intensity of

the basal reflection from the silicate layers. When insertion of polymer

chains in the silicate layers occurs, an increase of silicate interlayer

volume and corresponding layer spacing could be obtained which in turn

gives rise to the shifting of diffraction peaks to lower angles. Diffraction

peak cannot be seen in the case of exfoliated structures where silicate


EVA/Clay Nanocomposites Transport Features 231

layers are completely and uniformly dispersed in a continuous polymer

matrix [26] The X-ray diffraction patterns of the nano clay and polymer

nanocomposites are shown in Figure 9.1. Closite Na+ clay exhibits a

single peak at an angle 2 of 7o corresponding to a d-spacing of 11.7 Ao.

For EVA/clay nanocomposites, the characteristic diffraction peak moves to

a lower angle with respect to that of nano clay. For the composite samples

containing 3 (F3), 5 (F5) and 7 (F7) wt% of clay, the d-spacing were found

to be 16.3, 16 and 14.6 Ao corresponding to 2 5.4, 5.5 and 6.04o

respectively. This shows that EVA chains have intercalated into the

interlayers of closite Na+. It is found that in all systems the interlayer

spacing increases due to the intercalation of polymer into the layers of

nanoclay. Enhanced interlayer distance indicates that the layered structure

is retained. With the increase of clay content, the left shift magnitude of

diffraction peak decreases, that is, the enlargement extent of the interlayer

distance of the clay decreases. This indicates that lower the loading of

nanoclay the more favourable it is for the intercalation of EVA chains into

the silicate layers.


232 Chapter 9

+
Closite Na

Intensity

2
Intensity

F7

F5
F3

Figure 9.1 : XRD of nanoclay and EVA clay nanocomposites

9.2.1.2. Transmission electron microscopic analysis (TEM)

The transmission electron micrographs of various EVA-clay

nanocomposites are presented in Figure 9.2. The dark lines in the

transmission electron micrographs show the dispersion of silicates in the

polymer matrix. It can be seen that in F3 sample, the clay is well

dispersed in the matrix and is having a more ordered exfoliated structure.

When the percentage of clay increases, dispersion decreases and clay

exists as large aggregates and is unable to undergo exfoliation. The

above observation is consistent with the data observed from the XRD

patterns given in Figure 9.1.


EVA/Clay Nanocomposites Transport Features 233

a) EVA+DCP+3%F (F3) b) EVA+DCP+5%F (F5) c) EVA+DCP+7%F (F7)

Figure 9.2 : TEM images of nanocomposites

9.2.2. Positron annihilation lifetime spectroscopic analysis (PALS)

Free volume present in nanocomposite systems play a major role in

determining the overall performance of the membranes. PALS is an

efficient technique used for analysis of free volume. The diffusion of

permeant through polymeric membranes can be described by two

theories, viz. molecular and free volume theories. According to free

volume theory the diffusion is not a thermally activated process as in

molecular model but it is assumed to be the result of random

redistributions of free volume voids within a polymer matrix. Cohen and

Turnbull [27] developed the free volume models that describe diffusion

process when a molecule moves into void larger than a critical size; Vc.

Voids are formed during the statistical redistribution of free volume within

the polymer. The effect of layered silicates on o-Ps lifetime (3), o-Ps
234 Chapter 9

intensity (I3%) and relative fractional free volume % which are presented

in Table 9.1. It can be deduced from the table that relative fractional free

volume % is lowest for F3 system. It is found that the relative fractional

free volume of unfilled polymer decreases upon the addition of layered

silicates. The decrease is attributed to the interaction between layered

silicates and polymer due to the platelet structure and high aspect ratio of

layered silicates. The decrease is explained to the restricted mobility of

the chain segments in the presence of layered silicates. This results in

reduced free volume concentration or relative fractional free volume. The

contact surface area between the filler and the matrix is higher in

nanocomposites owing to its high aspect ratio, which in turn reduces the

free volume concentration. It is also found that the relative fractional free

volume % increases with clay loading. The increase in the values of

fractional free volume values, can be attributed to the aggregation of fillers

and the consequent additional void formation. The impact of nano particles

on the free volume and the barrier properties has been studied by Wang et

al. [3] and R. Stephen et al. [2]. They concluded that the permeability of

nanocomposite is mainly influenced by fractional free volume effects.


EVA/Clay Nanocomposites Transport Features 235

Table 9.1 : PALS measurement data of nanocomposites

o. Ps lifetime, o. Ps intensity Relative fractional


Sample
3 + ns I3 + 0.1% free volume %

F0 2.33 20.48 4.78

F3 2.38 7.29 1.77

F5 2.35 14.61 3.46

F7 2.32 17.85 4.13

9.2.3. Influence of nano particles on diffusion

The influence of nano clay on the sorption behaviour of EVA is presented in

Figure 9.3. The experiment was conducted at 28oC and the solvent used was

benzene. Unfilled (Fo) sample showed the maximum and filled sample with

3wt% of clay (F3) showed the least solvent uptake values. A similar trend

was also observed with other solvents. The increase in the barrier properties

of EVA membranes reinforced with layered silicates are due to the exfoliation

of silicates in the polymer matrix. The molecular level interaction of

polymer/clay results in reduced availability of free volume which in turn

reduces the diffusion through the membrane. The reduced sorption and

diffusion of filled membranes is owing to its platelet like morphology and high

aspect ratio of clay particles. Similar results were reported previously [2,15].

The influence of weight % of clay on the equilibrium uptake of solvents is

given in Figure 9.4. It is observed that, as the amount of clay increases, the

equilibrium uptake decreases. This is attributed to the difference in the

dispersion of clay particles in the matrix. TEM images shown in Figure 9.2
236 Chapter 9

clearly reveals that at higher filler loading, aggregation of filler particles occurs

due to its poor dispersion in the matrix. Thus a microphase separation was

formed between the polymer and clay particles, resulting in an increased

uptake of solvents. Similar results were reported in the literature [26].

F0
2.5
F3

2.0
Qt(mol%)

1.5

1.0

0.5

0.0
0 10 20 30 40
1/2 1/2
Time (min)

Figure 9.3 : Influence of nano particles on diffusion

2.6

2.5

2.4

2.3
Q

2.2

2.1

2.0

0 1 2 3 4 5 6 7 8

Filler loading (wt%)

Figure 9.4 : Influence of the weight percentage of filler on diffusion


EVA/Clay Nanocomposites Transport Features 237

9.2.4. Sorption behaviour

The sorption behaviour for the system under investigation has been

followed by the equation 3.1 (Chapter 3).

The values of n and k are determined by power regression analysis of the

linear portion of plots Qt versus square root of time. To ensure linearity,

values upto 50% of the equilibrium uptake were only used. The values of

n and k are placed in Table 9.2. The values of n for nanocomposites lie

in between 0.5 and 1, thus the sorption behaviour was found to be

anomalous.

Table 9.2 : Analysis of Sorption data at 28oC

n K x 10-2
Solvent
F0 F3 F5 F7 F0 F3 F5 F7

Benzene 0.59 0.67 0.66 0.64 2.72 1.8 1.98 2.10

Toluene 0.60 0.66 0.63 0.62 2.65 1.72 1.90 1.98

Xylene 0.57 0.67 0.64 0.63 2.36 1.63 1.89 1.93

9.2.5. Diffusion coefficient

The diffusion coefficient (D) was calculated using the equation 4.3

(Chapter 4).

The calculated values of diffusion coefficients are given in Table 9.2. It is

found that EVA/clay nanocomposite membranes show reduced diffusion

coefficient values. The result shows that the barrier properties of the

polymer nanocomposite membrane are remarkable. The values of


238 Chapter 9

diffusion coefficient show that the impermeable clay layers produce a

tortuous pathway for a penetrant to transverse the nanocomposite. This

not only enhances the barrier characteristics but also reduces the solvent

uptake. Chen et al. [28] showed the reason for the reduction of the solvent

diffusion coefficient of nanocomposites. They explained that the reduction

is due to the hindered diffusion pathways caused by the dispersion of the

individual nano sheets of the layered silicates in the nanocomposite.

The decrease in the diffusion rate of the polymer membranes modified with

nano clay is due to the nano metric level dispersion of the organic and

inorganic phases. Hence the available free volume decreases and this results

in the reduction of diffusion. Due to the platelet like morphology of silicates

the nano filled matrix exhibits reduced diffusivity owing to the increase in

tortuosity of the path. The ordered dispersion of clay is maximum for F3

sample, which is evident from TEM picture. Hence F3 sample showed the

least diffusivity but the diffusivity increases as a function of filler loading. This

can be explained in terms of aggregation of fillers at higher filler loading which

leads to an increase in free volume of the samples.

Table 9.3 : Values of diffusion coefficient (Dx1011m2s-1) at 28oc.

Solvent F0 F3 F5 F7

Benzene 3.1 1.54 1.89 2.68

Toluene 2.62 1.34 1.54 2.4

Xylene 2.1 1.26 1.42 1.67


EVA/Clay Nanocomposites Transport Features 239

9.2.6. Temperature effects and activation parameters

The temperature dependence of diffusion through nano clay filled EVA

was followed by conducting the experiments at 50 and 70oC in addition to

those at 28oC. In Figure 9.5, Qt mol% uptake is plotted as a function of

time at various temperatures for the F3 system. The solvent used was

benzene. It has been observed that maximum solvent uptake increases

with increase in temperature. All other systems showed the same trend. It

is also found that the slope of the linear portion increases with temperature

showing that the transport process is temperature activated.

6.5
6.0
5.5
5.0
0
4.5 28 C
0
4.0 50 C
Qt(mol%)

0
3.5 70 C
3.0
2.5
2.0
1.5
1.0
0.5
0.0
0 10 20 30 40
1/2 1/2
Time (min)

Figure 9.5 : Influence of temperature on the sorption behaviour of


nanocomposite

The energy of activation for the diffusion and permeation process is

calculated from the Arrhenius relationship.


240 Chapter 9

The values of activation energy for diffusion, ED and the activation energy

for permeation, EP were estimated. From the difference between EP and

ED, the heat of sorption, HS was estimated. The values of EP, ED, and

HS in benzene are complied in Table 9.4. It is found that activation

energy for diffusion and permeation for the unfilled sample is lower than

that of nano clay filled polymer membranes. The permeant molecules

require greater activation energy to travel through the layered silicates.

The large aspect ratio of the clay platelets, effectively increased the

diffusion path, which was responsible for the increased activation energy.

Table 9.4 : Activation parameters of diffusion

Sample ED kJ mol-1 EP kJ mol-1 HS kJ mol-1

F0 3.75 19.74 15.9

F3 19.6 33.28 13.68

F5 18.13 30.25 12.12

F7 14.8 26.4 11.6

The value of HS gives additional information about the molecular

transport through the polymer matrix. HS is a composite parameter

involving contribution from Henrys law and Langmuir type sorption. All

values are positive suggesting that, sorption is mainly dominated by

Henrys law, i.e. the formation of sites and the filling of these sites by

penetrant molecules.
EVA/Clay Nanocomposites Transport Features 241

9.2.7. Polymer-Solvent interaction parameter

The polymer-solvent interaction parameter () has been calculated from

the equation 4.11 (Chapter 4).

The polymer-solvent interaction parameter has been utilized to explain the

interaction between the solvents and the EVA samples. A low value of

indicates stronger interaction with solvents. The calculated values are

placed in Table 9.5. The values of the nano clay modified EVA samples

are higher than that of the unfilled sample. This shows that the interaction

of nanocomposites with the solvents is minimum. All the samples showed

maximum interaction with benzene.

Table 9.5 : Values of interaction parameter

Solvent F0 F3 F5 F7

Benzene 0.54 0.63 0.60 0.56

Toluene 0.58 0.68 0.66 0.62

Xylene 0.633 0.72 0.69 0.66

9.2.8. Network structure analysis

The molecular mass between crosslinks was estimated using the Flory-

Rehner equation 4.13 (Chapter 4).

The calculated MC values are given in the Table 9.6. The decrease in Mc

values of filled samples compared to unfilled one is due to the

reinforcement of clay in the polymer and hence the stiffness of the material
242 Chapter 9

increases. Lower values of Mc indicate that the network is more restrained

and this result in lower swelling of these samples.

Table 9.6 : Values of molecular weight between crosslinks (g/cc)

Mc Solvent F0 F3 F5 F7

Benzene 17291 9787 10264 12403

Mc Toluene 16412 7537 9816 10300

Xylene 14829 6994 7369 9968

The molecular weight between crosslinks (Mc) for the affine limit of the

model [Mc (aff)] was calculated using the equation 4.14 (Chapter 4). The

molecular weight between crosslinks for the phantom limit of the model [Mc

(ph)] was calculated using the equation 4.15 (Chapter 4). The values are

given in Table 9.7.

Table 9.7 : Values of molecular weight between crosslinks (g/cc)

Mc Solvent F3 F5 F7

Benzene 9595 9934 11854

Mc (aff) Toluene 7386 9619 10098

Xylene 6854 7074 9527

Benzene 4798 4967 5927

Mc (ph) Toluene 3693 4809 5049

Xylene 3827 3537 4763


EVA/Clay Nanocomposites Transport Features 243

It is found that Mc values of EVA/clay nanocomposites are close to Mc (aff).

This shows that in the solvent swollen state, the network deforms affinely.

9.2.9. Comparison with theory

The theoretical sorption curves were generated using the equation 3.6

(Chapter 3). Experimentally obtained values of diffusion coefficients are

substituted in the equation and the resulting curve is shown in Figure 9.6.

The theoretical and experimental results were not in good agreement. The

experimental curve deviates from the theoretical curve which is fully a

Fickian mode of diffusion.

1.0

0.8

Theoretical
0.6
Experimental
Qt/Q

0.4

0.2

0.0
0 5 10 15 20 25 30 35 40
1/2 1/2
Time (min)

Figure 9.6 : Comparison between experimental and theoretical sorption


curves of F3 at 28oC
B) Pervaporation characteristics of EVA/clay nanocomposites

9.2.10. Swelling ratio

Figure 9.7 shows the swelling ratio values of unfilled, and nano clay filled

EVA films. The unfilled membranes (F0) showed maximum swelling ratio
244 Chapter 9

values for all the feed concentrations. Modified EVA films with 3 wt% of

nano clay (F3) showed the minimum and the value increases with increase

in wt% of the filler.

The two main factors that influence the swelling of the films; the fraction of

the amorphous phase in the polymer and the chemical compatibility

between the polymer chain and the solvent mixture. When the crystalline

fraction of the polymer decreases there is an increase of both the volume of

amorphous fraction and chain lengths that connect the crystalline domains.

A higher material volume accessible for the liquid sorption and a higher

flexibility of the network allow an increased solvent uptake.

5.0
F0
4.5 F3
4.0 F5
F7
3.5
Swelling ratio

3.0

2.5

2.0

1.5

1.0

0.5

0.0
20 30 40 50 60 70 80 90

Chloroform in feed (wt%)

Figure 9.7 : Swelling ratio of unfilled and EVA nanocomposite films

The decreased swelling ratio values of filled EVA films (F3) is explained

as follows. The impermeable clay layers dictate a tortuous path way


EVA/Clay Nanocomposites Transport Features 245

for a permeate to pass through the nanocomposite. The diffusion path

is schematically represented in Figure 9.8.

(a)

(b)
Figure 9.8 : Schematic representation of diffusion through
(a) composite with conventional filler (b) nanocomposites

From the Figure 9.8 (a), it is clear that as in the case of micro

composite the penetrant molecules can easily pass through the


246 Chapter 9

interphase between filler and the matrix. However, in the case of clay

filled nanocomposite (Figure 9.8 (b)) penetrant molecule experiences a

difficult pathway due to molecular level dispersion of clay in the matrix.

9.2.11. Pervaporation of chloroform-acetone mixtures

The pervaporation performance of unfilled and EVA nanocomposite

membranes were analysed using chloroform-acetone mixtures. Both

unfilled and nano clay filled membranes showed chloroform selectivity

from chloroform-acetone mixtures. The affinity of EVA membranes

towards chloroform is higher than acetone and this creates a remarkable

difference in the separation of chloroform from chloroform-acetone

mixtures [29].

Table 9.8 shows the permeation rate and the selectivity of unfilled

(F 0) and modified films with 3wt% of filler (F3) membranes. EVA/ clay

nanocomposites showed a higher selectivity but a lower permeation

rate than the unfilled ones. The increased selectivity is due to the

exfoliation of silicates in the polymer matrix leading to the nanometric

level dispersion of the organic and inorganic phases. The molecular

level of polymer/filler interaction results in a reduced availability of

free volume, as a result the permeation rate decreases and

separation factor increases. The enhanced selectivity of nano filled

membranes are owing to its platelet like morphology and high aspect

ratio of the fillers. Due to the high aspect ratio of layered silicates

the contact area between filler and the matrix increases. Hence
EVA/Clay Nanocomposites Transport Features 247

there will be more resistance towards molecular diffusion resulting in a

reduced permeation rate. The above results have been complemented by

PALS analysis.

Table 9. 8 : Pervaporation characteristics of EVA films (22 wt% of


chloroform-acetone mixture)

Permeation
2
System composition Selectivity ij Flux (kg/m h)
(wt%)

F0 52 3.8 0.27

F3 91 36 0.12

9.2.12. Influence of nano clay loading

The influence of the wt% of nano clay on pervaporation performance is given in

Figure 9.9. The selectivity factor decreases sharply when the clay content

becomes higher than 3 wt%. This can be explained in terms of aggregation of

clay particles with increase in concentration of clay, resulting in the weakening

of polymer chain. When the clay composition is greater than 3 wt%, the

compatibility of filler and the matrix decreases, resulting in microphase

separation. PALS analysis also showed that the fractional free volume %

decreases at higher clay loading. Hence a drop in selectivity and an increase

in permeation rate was observed. Wang et al. [26] investigated the effect of

clay content on the pervaporation performance of 90wt% ethanol aqueous

solution through the polyamide/clay nanocomposite membrane. They


248 Chapter 9

found that selectivity decreases sharply when the clay content becomes

higher than 2wt%.

0.22
20
Selectivity
Flux

Permeation flux (kg/m .h)


18

2
0.20

16
Selectivity

14
0.18

12

10 0.16

6 0.14
3 4 5 6 7

Clay loading (wt%)

Figure 9. 9 : Influence of filler loading on pervaporation

9.2.13. Calculation of membrane selectivity ( mem)

The overall selectivity is a combination of the effects of the membrane

selectivity (sorption-diffusion selectivity) and the volatility or evaporative

selectivity. If the downstream pressure is negligible, the apparent

separation factor or selectivity is given by [30].

ij = mem. evp (9.2)

( Pi ) ri . pi
= (9.3)
( Pj ) r j . p j
EVA/Clay Nanocomposites Transport Features 249

where P refers to the permeability and p and r refers to vapour pressure

and activity coefficients of the components i & j. The membrane selectivity

( mem) can be calculated using the formula.

ij
mem = (9.4)
evp

The membrane selectivity, mem is calculated for unfilled (F0) and

nanocomposite membrane (F3) using the equation (9.4). The values of

membrane selectivity will provide an interesting study of how a mass

separating agent can overcome the intrinsic volatility differences and

enables to permeate the less volatile component in a mixture. Table 9.9

gives the membrane selectivity for unfilled and nanocomposite

membrane. It can be seen from the table that the evaporative selectivity is

being overcome by the membrane selectivity. The membrane selectivity is

much higher for nano clay modified EVA membranes.

Table 9.9 : Membrane performance (22 wt% of chloroform mixture)

Sample mem ev

F3 72.5 0.49

F0 6.6 0.49
250 Chapter 9

9.2.14. Comparison of pervaporation results with vapour-liquid


equilibrium (VLE) data
Chloroform and acetone form an azeotrope at 80 wt% of chloroform.

Separation of azeotropes by simple distillation is possible only by adding a

third component, i.e., an entrainer such as benzene, which is known as

deadly carcinogen. In the membrane based pervaporation separation,

membrane acts as a third phase to break the azeotrope. Thus

pervaporation is more effective in separating azeotropes than conventional

distillation. Figure 9.10 shows that pervaporation curve is higher than that

of the vapour-liquid equilibrium (VLE) curve throughout the composition

range of the feed mixture.

1.0

0.8
Chloroform fraction(w/w vapour)

0.6
B
D

0.4

0.2

0.0
0.0 0.2 0.4 0.6 0.8 1.0

Chloroform fraction(w/w feed)


B:Vapor composition of the permeate
D:Liquid-vapour equlibrium curve

Figure 9. 10: Comparison of vapour-liquid equilibrium curve with


pervaporation data for nanocomposite films.
EVA/Clay Nanocomposites Transport Features 251

C) Transport of organic vapours through EVA / clay


nanocomposites
9.2.15. Mechanical properties

The tensile properties of EVA-clay nanocomposites are shown in Table 9.10.

The tensile strength of the EVA-clay composites are higher than the unfilled

ones. The tensile strength of EVA/clay nanocomposite with 3 wt% clay is

25% higher than that for the EVA membrane. This may be due to the

reinforcing effect of the silicate sheets through the random dispersion in

the polymer matrix. However, the tensile strength decreased on further

increase in clay content and it is reduced to 6.8 MPa for samples

containing 7wt% of clay content. The decrease in tensile strength at higher

filler loading is due to the uneven distribution of stress by the aggregated

clay present in the matrix.

Table 9. 10 : Effect of clay content on tensile properties

Clay Content (Wt%) Tensile Strength (MPa)

0 6.2

3 7.8

5 7.7

7 6.8

9.2.16. Vapour transport

The vapour permeation coefficient of unfilled and nano clay modified

composites are shown in Table 9.11. The vapour permeation

characteristics of EVA/clay nanocomposites were analysed using


252 Chapter 9

chloroform. The value of permeation coefficient is maximum for unfilled

membranes. Among the filled samples, the permeation coefficient

increases with increase in weight percentage of clay. The reduced vapour

permeability of membranes reinforced with layered silicates are due to the

molecular level of polymer / clay interaction resulting in a reduced free

volume. The large aspect ratio of the clay platelets effectively increase the

penetration path, which was responsible for the reduced permeability.

However, it was found that when the filler weight percentage was greater

than 3%, there is a gradual increase in permeability due to the decrease in

the dispersion of nano particles.

Table 9. 11 : Permeability coefficient [Px1010 (mol Pa)]

System Permeability

F0 7.6

F3 2.9

F5 3.7

F7 4.2

D) Gas transport through EVA/clay nanocomposites

9.2.17. EVA/Clay nanocomposite membranes for gas separation

The gas transport properties of nano clay reinforced polymer membranes have

been analysed using oxygen and nitrogen gases. The results were compared

with that of unfilled ones (F0). Oxygen and nitrogen gas permeability

coefficients are shown in Figures 9.11 and 9.12 respectively. It is found


EVA/Clay Nanocomposites Transport Features 253

that the transport of gases through layered silicate filled membranes is

lower than that of unfilled ones. The enhancement in gas barrier properties

of nano clay reinforced membranes indicate strong polymer/filler

interaction. Since the chain segments get immobilized in the presence of

layered silicates, the free volume decreases and as a result the gas

permeability coefficient reduces. Utracki and co-workers [31] studied the

reduced free volume available in the polymer matrix after the incorporation

of clay platelets. According to them, in exfoliated polymer nanocomposite

the accessible clay surface area is proportional to organo clay loading.

They observed that the addition of 4 wt% of organo clay (closite 15) can

reduce the matrix hole fraction twice as large as that observed for polymer

nanocomposite with 2wt %. The incorporation of 1.1 and 2.42 wt% of

montmorillonite (MMT) can reduce the matrix free volume to 4.7 and 8.0%

respectively.

5
1bar
O2gas
4
P x 10 (mol/msPa)

3
10

0
F0 F3 F5 F7

Figure 9. 11 : Variation in oxygen permeability


254 Chapter 9

2.2
1bar
2.0
N2gas
1.8

1.6
P x 10 (mol/msPa)

1.4

1.2

1.0
10

0.8

0.6

0.4

0.2

0.0
F0 F3 F5 F7

Figure 9. 12 : Variation in N2 permeability

The addition of fillers reduce gas permeability of polymers according to a

tortuous path model developed by Neilson [32].

1- Q f
Pc = Pp (9.5)
1+ Qf /2

where Pc and Pp are the permeability of composite and polymer, Qf is the

volume fraction of filler and is the aspect ratio of platelets. The

calculated value of is 112.9. The high aspect ratio of the clay platelets

effectively increased the gas penetration path, which is responsible for the

reduced permeability. Schematic representation of the tortuous path model

is given in Figure 9.13. From the figure, it is clear that the gas molecules

have to travel through a tortuous path in the presence of layered silicate.

The reduced gas permeability of nanocomposites is influenced by two

factors, viz. geometry of the filler and the molecular level interaction of the
EVA/Clay Nanocomposites Transport Features 255

matrix and the clay. Also the extent of exfoliation is found to be maximum

in F3 samples (Figure 9.2 (a)). Therefore, these samples exhibit reduced

gas permeability.

Figure 9.13: Schematic representation of the tortuous path model


developed by Nielsen for the transport of gases through
filled membranes

9.2.18. Selectivity of membranes

The polymeric membranes used for gas separation processes have certain

significance such as high permeability to the desired gas, high selectivity

and the ability to form useful membrane configurations. The requirement

of an ideal membrane is high permeability along with high permselectivity.

The permselectivity of membrane is given by

P(O2 )
(O2, N2) = (9.6)
P(N2 )
where, is the permselectivity of a membrane towards O2 and N2 gas,

P(O2) and P(N2) are the permeability coefficients of O2 and N2 gases,

respectively.
256 Chapter 9

The permselectivity values of the membranes are given in Table 9.12. The

nanocomposites possess higher selectivity than the unfilled one. These

might be ascribed to the interaction between the polymer and the filler.

Table 9. 12: Oxygen to Nitrogen selectivity values of unfilled and filled


EVA membranes [P(O2)/P(N2)]

Sample Permselectivity

F0 3.5

F3 5.02

F5 4.5

F7 3.8

9.2.19. Effect of pressure

The permeation mechanism can be obtained by examining the variation in

permeability coefficient as a function of pressure. The effect of pressure on

nitrogen gas permeability of F0 and F3 membranes is given in Figure 9.14.

The unfilled (F0) system exhibit increase in permeability with pressure. This is

due to the higher solubility of the permeant molecules in the polymer chain as

a function of pressure. The effect of pressure on the filled system is negligible

due to the close packing of fillers in the polymeric matrix.


EVA/Clay Nanocomposites Transport Features 257

1.6
F0
1.4 F3

1.2
Px10 (mol/m.sPa))

1.0

0.8
10

0.6

0.4

0.2
0.4 0.6 0.8 1.0 1.2 1.4 1.6

Pressure(bar)

Figure 9. 14 : The effect of pressure on the permeability of nitrogen gas


through filled (F3) and unfilled membranes

9.3. Conclusion

EVA/clay nanocomposites containing different filler loading have been

prepared and the transport features of the membranes were

investigated. Morphology of the composite membranes were analysed

by XRD and TEM. It has been found that the diffraction peaks were

shifted to lower angles with an increase in d-spacing. Samples with

3wt% of filler showed maximum increase in d-spacing. TEM images

showed that sample with 3 wt% of clay showed excellent dispersion of

clay particles resulting in an exfoliated structure. The dispersion of

nano particles decrease with an increase in the clay loading. The

fractional free volume % was determined using positron annihilation

lifetime spectroscopic analysis. Sample with 3 wt% clay showed the

least free volume.


258 Chapter 9

The liquid transport characteristics of EVA/clay nanocomposite

membranes were investigated using aromatic hydrocarbons as probe

molecules. Due to enhanced polymer/filler interaction, sample with 3wt%

of clay exhibited lower solvent uptake. However, the solvent uptake

tendency increased at higher clay loading. This is ascribed to the poor

physical interaction between the matrix and filler, leading to aggregation

of fillers. The diffusion coefficient values also showed the same trend.

The mechanism of transport was found to be anomalous. The

pervaporation performance was analysed using chloroform-acetone

mixtures. These membranes exhibited a far superior selectivity to

chloroform molecules than the unfilled one. The vapour permeability was

examined using chloroform vapours and nanocomposite membranes

exhibited very low vapour permeability compared to unfilled one.

The gas transport properties of nano clay filled and unfilled (F0)

membranes were investigated using permeant gases such as O2 and

N2. Due to enhanced polymer/filler interaction, nanocomposite

membranes exhibited lower permeability to oxygen and nitrogen gases.

Increase in effective penetration path due to the very large aspect ratio

of the silicate layers was responsible for the reduced gas permeability.

As the filler loading increased the permeability of the polymer increased

due to the aggregation of the filler particles. From the plots of pressure

against permeability it can seen that pressure has little influence on the

permeation of gases through nanofilled composites. Finally it is


EVA/Clay Nanocomposites Transport Features 259

important to mention that by the incorporation of nanofillers into EVA

matrix, new gas barrier membranes could be developed.


260 Chapter 9

Reference

[1] S. Takahashi and D. R. Paul, Polymer, 47, 7535, (2006).

[2] R. Stephen, S. Varghese, K. Joseph, Z. Oommen and S. Thomas,

J. Membr. Sci., 282, 162 (2006).

[3] Z. F. Wang, B. Wang, N. Qi, H.F. Zhang and L. Q. Zhung, Polymer,

46, 719 (2005).

[4] J. K. Kim, C. Hu. R. S. C. Woo, and M. L. Sham, Comp. Sci.

Technol., 65, 805 (2005).

[5] Q. Liu and, D. De Lee, J. Non-Newtonian Fluid Mech., 131, 32

(2005).

[6] T. Jiang, Yu-huawang, J. Yeh and Zhi-giang Fan, Eur. Polym. J., 4,

459 (2005).

[7] S. Wang, C. Long, X. Wang, LiQ, and Qiz, J. Appl. Polym. Sci., 69,

1557 (1998).

[8] D. C. Lee and L. W. Jang, J. Appl. Polym. Sci., 68, 1997, (1998).

[9] M. W. Noh and D. L. Lee, Polym. Bull., 42, 619, (1999).

[10] Y. Yang, Z. K. Zhu, J. Yin, X. Y. Wang and Z-e Qi, Polymer, 40,

4407 (1999)

[11] J. J. Luo and I. M. Daniel, Compos. Sci. Technol., 63, 1607 (2003).

[12] M. Krook, A.C. Albertsson, U.W. Gedde, M.S. Hedenqvist, Polym.

Eng. Sci., 42, 1238, (2002).


EVA/Clay Nanocomposites Transport Features 261

[13] R. K. Bhardwaj, Macromolecules, 34, 9189, (2001).

[14] A. D. Drozdov, J.D. Christiansen, R.K. Gupta and A.P. Shah, J.

Polym. Sci. Part. B : Polym. Phys., 41, 476 (2003) .

[15] M.S. Hedengvist, A. Backman, M. Gallstedt, R.H. Boyd and U.W.

Gedde, Comp. Sci. Technol., 241, 156, (2006).

[16] P. Musto, L. Mascia, G. Mensiteri and Rayosta, Polymer, 46, 4492

(2005)

[17] T. C. Merkel, Z. He, Pinnaul, B.D. Freeman, P. Meakin and A.J. Hill,

Macromolecules, 36, 6844 (2003)

[18] D. P. N. Vlasveld, J. Groenewold, H. E. N. Bersee and S. J. Picken,

Polymer, 46, 12567 (2005).

[19] T. A. Shantalii, I. L. Karpova, K. S. Dragan, E. G. Privalko and

UV.P. Privalko, Sci. and Tech. of Advanced Materials, 4, 115

(2003).

[20] M. Sairam, B. V. K. Naidu, S. K. Natraz, B. Sreedhar and T. M.

Aminabhavi, J. Membr. Sci., 283, 65 (2006).

[21] A. Okada, M. Kawasumi. M,. Unuki. A, Kojima. Y, Kuranchi. T,

Kamigailo. O, Mater. Res. Sac. Prox., 45, 171 (1990).

[22] S. Varghese and J. Karger-Kocsis, Polymer, 44, 4921 (2003).

[23] B. Barbi, S. S. Funari, R. Gehrke, N. Scharnagl, N. Stribeck,

Macromolecules, 36, 749 (2003).


262 Chapter 9

[24] X. Fu and S. Qutubuddin, Polymer, 42, 807 (2001).

[25] N. Hasegawa, H. Okamoto, M. Kato, A. Usuki, N. Sato, Polymer,

44, 2933 (2003).

[26] Y.C. Wang, S.C. Fan, K.R. Lee, C. L. Li, S.H. Huang, H.A. Tsai and

J.Y. Lai, J. Membr. Sci., 239, 219 (2004).

[27] D. Turnbull and M.H. Cohen, J. Chem. Phys., 34, 120 (1961).

[28] C. Chen, M. Khobaih and D. Curliss, Progress in Organic Coatings,

47, 376, (2003).

[29] S. Anilkumar, P.H. Gedam, V.S. Kishan Prasad, M.G. Kumaran and

Sabu Thomas, J. Appl. Polym. Sci., 60, 735, (1996).

[30] Wenyuan Zil, Donald R. Paul, William J. Kores, J. Membr. Sci., 89,

219 (2003).

[31] L. A. Utracki and R. Simha, Macromolecules, 37, 10123 (2004).

[32] L. E. Neilson, J. Macromol Sci. (Chem)., A1 (5), 929 (1967).

Вам также может понравиться