Sei sulla pagina 1di 7

Polymer Degradation and Stability 63 (1999) 391397

Eect of interactions between stabilisers and silica used for antiblocking applications on UV and thermal stability of polyolen lm
1. Adsorption studies
C.M. Liauw a,*, A. Childs a, N.S. Allen a, M. Edge a, K.R. Franklin b, D.G. Collopy c
a

Department of Chemistry and Materials, The Manchester Metropolitan University, Chester Street, Manchester M1 5GD, UK
b
Unilever Research, Port Sunlight Laboratory, Quarry Road East, Bebington, Wirral L63 3JW, UK
c
Croseld Limited, Research and Development, PO Box 26, Warrington, Cheshire WA5 1AB, UK
Received 1 May 1998; accepted 7 June 1998

Abstract
Interactions between stabilisers and silicas, produced by both the gel process and by precipitation, have been studied using ow
micro-calorimetry (FMC) and diuse reectance Fourier transform infrared spectroscopy (DRIFTS). The silicas had approximately equivalent BET surface areas. By virtue of its greater pore volume and lower water content, the gel silica was shown to have
the greatest adsorption activity. For both the silicas, the adsorption activity of a stabiliser was related to its basicity and number of
adsorbing groups per molecule which gave rise to multi-point adsorption, hence polymeric hindered amine light stabilisers (HALS)
and large hindered phenols showed the strongest adsorption. Competitive adsorption of polymeric HALS versus hindered phenols
was also investigated. The data obtained has been very useful for predicting preferential adsorption of the stabilisers on to the
silicas from a linear low density polyethylene melt which in turn has been related to photo-oxidative and thermo-oxidative stability.
The latter investigation will be reported separately. # 1999 Elsevier Science Ltd. All rights reserved.

1. Introduction
In production of layat lm, self adhesion (blocking)
of the lm layers after bubble collapse is a serious problem which can be overcome by addition of very low
levels (i.e. 10002000 ppm) of a suitable ller to the
polymer, such additives are known as antiblocking
agents [1]. It is anticipated that the ller particles serve
to roughen the surface of the lm and hence reduce the
contact area thereby reducing the level of self adhesion
[1]. Antiblocking llers include calcium carbonate and
silica, which is the subject of this study.
The stabilisation of polymer lms is an important
issue due to their high surface area to volume ratio
which causes the mechanical properties to be highly
surface dependent. Adsorption of polymer additives on
to llers is an acknowledged problem, particularly in the
elastomer industry where measures have to be taken
against such eects [2]. In thermoplastics, the adsorption of stabilisers by llers is a recognised problem
which is solved by using additives which sacricially
* Corresponding author. Tel.: +44-161-247-3325; fax: +44-161247-6357; e-mail: c.m.liauw@mmu.ac.uk

adsorb on to ller surfaces thereby blocking adsorption


of the stabilisers [3]. Previous work by the authors [4,5],
however, has indicated that adsorption of stabilisers by
high surface area silicas is not necessarily a negative
feature and can be used to provide a what may be considered a reservoir of stabiliser [4]. Along with sacricial
adsorption eects [5], interactions with the silica surface
may ensure that the additives are released slowly.
Therefore it is possible that this eect prevents leaching
and volatilisation of the stabilisers thus aording longer
lasting protection of the polymer.
In order to fully understand such controlled release
behaviour, the interactions between stabilisers and ller
must be well understood. Flow micro-calorimetry has
proved to be a useful tool for studying the interactions
between llers and surface modier/polymer additive
molecules [68]. FMC can determine both heat of
adsorption/desorption and the amount adsorbed/desorbed and is also useful for the study of competitive
adsorption.
Amorphous silica is a complex substrate due to variation in SiO bond angles at the surface which leads to
varying distances between silanol groups and thus results
in silanol groups of varying activity [9]. Decreasing the

0141-3910/99/$see front matter # 1999 Elsevier Science Ltd. All rights reserved
PII: S0141 -3 910(98)00119 -0

392

C.M. Liauw et al./Polymer Degradation and Stability 63 (1999) 391397

distance between silanol groups increases the possibility


of mutual hydrogen bonding. It can be envisaged that
the radius of surface curvature, and whether this is
concave or convex, has a signicant eect on the
separation distance of the silanols. Gel silicas have an
extensive pore structure in which silanol activity on the
pore walls varies according to pore radius. With pore
diameters less than 10 nm silanols become close enough
to allow a large fraction of the silanols to be mutually
hydrogen bonded [10]. The silanol groups on the surfaces of pores greater than 10 nm are similar to those on
a plane surface and hence are not likely to be mutually
hydrogen bonded.
Isolated silanol groups are considered to be more
reactive towards polar and aromatic species [11],
whereas mutually hydrogen bonded silanols only tend
to interact with strong hydrogen bonding donors such
as water and methanol [12]. These ndings were more
recently conrmed by Mauss and Englehardt [13] who
used diuse reectance Fourier transform infrared
spectroscopy (DRIFTS) to follow the eect of thermal
treatment of a chromatographic gel silica. Interaction of
the silica with alcohols, ketones and esters was assessed
chromatographically and related to the proportion of
isolated to vicinal (mutually hydrogen bonded) silanols.
Loss of vicinal silanols, accompanied by an increased
level of isolated silanols, between 400 and 600 C, resulted
in signicantly decreased interaction with alochols and
slightly increased interaction with ketones and esters.
Adsorption of atmospheric moisture on to the silanol
groups [14] complicates the above picture, the rst layer
of water is adsorbed with oxygen facing the surface;
further molecules can adsorb onto the rst layer creating hydrogen bonded clusters around adjacent pairs of
isolated silanols [15]. Water adsorbed via hydrogen
bonding on to isolated silanols is highly dissociated and
is able to open the siloxane bond and thus create more
silanols [16]. It is therefore evident that adsorbed water
can have a signicant eect on the surface chemistry of
silica and hence its adsorption activity. However, it is
very dicult to remove all the water from silica without
removing some of the silanol groups.
Another variable that can eect adsorption behaviour
is the presence of metal ion impurities within the silica
structure. Aluminium ions on the silica surface form
Lewis acid sites and transition metal ions, such as iron,
can complex with ligands such as amines. Sodium ions
(usually associated with sodium oxide) interact very
strongly with water which can result in retention of
water and silanol groups within the silica structure [17].
In this study, interactions between a range of UV
stabilisers/antioxidants and two silica samples were
investigated using FMC. Competitive adsorption
between selected UV stabilisers and antioxidants was
also investigated. Mode of adsorption was investigated
using diuse reectance Fourier transform infrared

spectroscopy. Eects of these interactions on the stabilisation performance in linear low density polyethylene
will be reported separately.
2. Experimental
The stabilisers investigated were Irganox 1010, Chimassorb 944, Tinuvin 328 [all from Ciba, UK] and BHT
(2,6 Di-tert-butyl-4-methylphenol) (Aldrich), further
details are given in Table 1 and the structures are presented in Fig. 1. All other solvents/reagents were of
HPLC grade or 99%+ purity where appropriate. The
silicas used were supplied by Croselds Limited, Silica A
was produced by the gel process and Silica B produced
by precipitation. Other parameters are given in Table 2.
The FMC system used was that developed by Ashton
and Rothon [5] and consisted of a Microscal 3V
(upgraded to all PTFE uid path) FMC linked to a
Waters 410 dierential refractometer. The data outputs
were handled by PerkinElmer Nelson 900 Series data
stations which were in turn linked to a PC running
Nelson chromatography software. Experiments were
conducted using a cell temperature of 27 (1 C) and
an additive concentration of 0.30% w/v. The n-heptane
carrier ow rate was 6.6 cm3 hr1. n-Decane was used as
the non-adsorbing probe. Sample size was 24 mg (1
mg) and 45 mg (1 mg) for Silicas A and B, respectively. Before adsorption experiments were commenced
the sample bed was allowed to equilibrate for two hours
at a carrier ow rate of 6.6 cm3 hr1. Three separate
runs were carried out with each additive and the average
results reported. Competitive adsorption was investigated by rst attaining adsorption equilibrium with one
stabiliser and then switching to a solution of the other in
the same solvent. Any adsorption/displacement was
observed in the usual manner. The experiment was then
repeated on a fresh bed of silica, this time with the order
of stabiliser addition reversed. For a more comprehensive description of the FMC technique, the reader is
referred to a description by Ashton and Briggs [18].
Samples isolated from the FMC were air dried and
then examined using DRIFTS. Before analysis the
samples were diluted to 5% w/w with nely ground KBr
and then placed in a Spectra-Tech DRIFTS cell tted to
a Nicolet 510 FTIR spectrophotometer (DTGS detector). Spectra were made up of 150 scans with resolution
set to 4 cm1. All spectra shown are normalised to a
silica band at 800 cm1 which is not aected by silanol
type/concentration [13].
3. Results and discussion
The FMC data is represented in Figs. 2 and 3. It is
immediately evident that Silica A has a higher overall

C.M. Liauw et al./Polymer Degradation and Stability 63 (1999) 391397

393

Table 1
Stabilisers investigated
Trade name
(supplier in parentheses)

Additive type

Systematic name

Molar mass

Irganox 1010 (Ciba)

AO

1178

BHT (Aldrich)
Chimassorb 944 (Ciba)

AO
LS

Tinuvin 328 (Ciba)

LS

Benzenepropanoic acid-3, 5-bis(1,1-dimethyl)-4-hydroxy-2,


2-bis[[3-[3,5-bis(1,1-dimethylethyl)-4-(hydroxyphenyl]-1-oxopropoxy]methyl]-1,
3-propanediylester
2,6-Di-tert-butyl-4-methyl phenol
Poly[[6-[(1,1,3,3-tetramethylbutyl)amino]-1,3,5-triazine-2,4-diyl][2,2,6,
6-tetramethyl-4-piperidinyl)imino]-1,
6-hexanediyl[2,2,6,6-tetramethylpiperidinyl) imino]]
2,20 -hydroxy-30 50 -di-tert.amylphenyl)-2H-benzotriazole

Fig. 1. Structures of stabilisers used, (a) Irganox 1010, (b) BHT, (c)
Chimassorb 944, (d) Tinuvin 328.

218
ca. 3000
335

adsorption activity, particularly with the stabilisers


which have potential for multi-point adsorption, i.e.
Chimassorb 944 and Irganox 1010. Of these two, Chimassorb 944 showed the largest heat of adsorption and
amount adsorbed. The smaller additives, i.e. BHT and
Tinuvin 328 have no scope for multi-point attachment
and are only weakly adsorbed. With Silica B the overall
adsorption activity was far less, but Chimassorb 944
and Irganox 1010 were still adsorbed to a greater extent,
however, in this case the dierence between the two was
not so marked. On Silica A the levels of adsorption were
very high with the result that only a small amount of
this ller (ca. 0.6% w/w) would be needed to consume
all the Chimassorb 944 (ca. 0.2% w/w) in a typical formulation. Earlier work [4,5] has indicated the potential
benets of such an eect.
The mode of adsorption of the additives is apparent
from the DRIFTS data (Fig. 4). For Silica A treated in
the FMC cell with Irganox 1010, the following spectral
perturbations are apparent. The ester carbonyl is broadened and shifted to lower energy [Fig. 4(b) and (c)],
relative to when the additive is unbound [Fig. 4(d)]. The
OH stretch (3627 cm1) of the Irganox 1010 is also
reduced in relative intensity. These observations clearly
reect adsorption via the ester groups and phenolic OH
group. Similar observations are also apparent for Chimassorb 944 where a reduction in NH stretching intensity was observed thereby indicating adsorption via the
substituted piperidine portion of the molecules. With
both these adsorbates, it is evident that after adsorption,
the isolated silanol absorption of the silica [3745 cm1,
Fig. 4(a)] disappears and hence conrms the participation of these species in the adsorption interactions.
It is appreciated that some of the additives examined
are large and complicated molecules with sometimes
several features that may promote adsorption. In order
to attempt to further unravel the mechanism of additive
adsorption, studies were conducted with a series of
model compounds, some of which represent a selected
portion of the additve molecule concerned. Relevant
data is presented in Figs. 5 and 6. It is again evident
that Silica A has the greater adsorption activity, being

394

C.M. Liauw et al./Polymer Degradation and Stability 63 (1999) 391397

Table 2
Properties of silicas A and B
Silica Surface areaa
(m2 g1)
A
B

320
364
a
b

Adsorption pore volumea


(cm3 g1)

Desorption pore volumea


(cm3 g1)

1.6
0.3

1.8
0.4

Water content Pore volume Iron content Na2O content


(%w/w)
(cm3 g1)b
(ppm)
(ppm)
5.8
11.4

1.5
0.7

23
215

595
1600

Obtained by N2 BET measurements.


Obtained by mercury porosimetry.

Fig. 2. FMC data for Silica A (adsorption data is all exothermic, the
sign of the energy change has been reversed to facilitate comparison):
& adsorption, & desorption.

Fig. 3. FMC data for Silica B (adsorption data is all exothermic, the
sign of the energy change has been reversed to facilitate comparison):
& adsorption, & desorption.

receptive to both oxygen containing and nitrogen containing model compounds, with the more basic amine
probes, i.e. ethylamine, diethylamine and the substituted
piperidine, having the higher heats of adsorption. Silica
B presents a completely dierent picture; in this case the
overall amount of probe adsorption, particularly with
ketones and esters, is less than for Silica A, however, the
energies of adsorption per molecule and strengths of
adsorption for the more basic amine probes are signicantly higher. This shows that Silica B may possess a
greater number of Brnsted acid sites relative to Silica A.
These observations generally mirror those for the stabilisers and conrm that adsorption can occur via hydroxyl
groups, amine functionality and ester linkages.

These inferences at rst seem surprising considering


the similarity in silica surface area, however, they can be
explained by the dierences in both pore size distribution and pore volume (Table 2) which in turn aects the
proportions of isolated, and mutually hydrogen bonded
silanol groups [11]. At this juncture it must also be
pointed out that the water and metal ion contents of the
silicas are dierent; Silica B has twice the water content
of A, a feature which was later conrmed by DRIFTS.
This water is not likely to be completely removed by the
heptane carrier uid. Water adsorbed on silica can be
highly dissociated, resulting in the formation of Brnsted
acid sites. This eect alone may explain the higher heats
of adsorption observed with the 1, 2, 3 amines and

C.M. Liauw et al./Polymer Degradation and Stability 63 (1999) 391397

Fig. 4. DRIFTS data for Silica A isolated from FMC cell: (a)
untreated silica; (b) Silica A after adsorption/desorption of Irganox
1010; (c) as (b) but substrate subtracted; (d) unbound Irganox 1010
(run as powder in KBr).

395

pyridine. Silica B also has a signicantly higher iron


content which again could explain the activity with the
above amines via chelation.
The range of silanols and adsorbed water can be differentiated using infrared spectroscopy [13,17]: isolated
silanols give rise to a sharp absorption at 3740 cm1 and
vicinal silanols absorb at 3600 cm1. Surface water,
which is hydrogen bonded results in a broad absorption
at 3400 cm1. DRIFTS analysis (Fig. 7) was carried out
on Silicas A and B after conditioning at 95 C in vacuo
for 16 h. This treatment should remove the majority of
the physically adsorbed water [9,19] and hence facilitate
resolution of the silanol groups. It is immediately evident that Silica B contains a higher level of strongly
bound (hydrogen bonded) water, some of which may be
trapped within the structure of the silica as a result of its
higher sodium content. Silica A has a higher ratio of
isolated to vicinal silanols than Silica B, an observation
which, along with its higher pore volume, may contribute

Fig. 5. FMC data for adsorption of model compounds: & adsorption, & desorption (4A2266T is 4-amino-2,2,6,6-tetramethylpiperidine).

396

C.M. Liauw et al./Polymer Degradation and Stability 63 (1999) 391397

reverse experiment in which adsorption of Chimassorb


944 is followed by addition of Irganox 1010, showed no
evidence of adsorption of the second probe. This was
subsequently conrmed by DRIFTS analysis on the
isolated ller which again showed total coverage with
Chimassorb 944.

to its greater overall adsorption activity. In the case of


Silica B some of the surface silanols may be blocked by
hydrogen bonded water which results in only weak
interaction with adsorbates that have a predominant
hydrogen bonding acceptor/Lewis base character.
However, the more basic probes (i.e. the amines) are
able to interact strongly with the Brnsted acid sites
produced via dissociation of the hydrogen bonded
water. Vicinal silanols on Silica B will also be able to
interact via hydrogen bonding with these amines and
phenol. Hydrogen bonded water on the silica surface
will reduce the concentration of Lewis acid sites (i.e.
isolated silanols) which also interact strongly with
ketones and esters as well as amines.
Competitive adsorption can also be readily studied by
FMC. Fig. 7 shows FMC data for adsorption of Irganox
1010 followed by Chimassorb 944 on to Silica A. Fig. 8
shows equivalent data for the reverse experiment. From
this data is was apparent, by virtue of evolution of solvent [Fig. 7(b)], that Chimassorb 944 is initially adsorbed at sites that are not occupied by Irganox 1010 which
is adsorbed to saturation level. Considering the slightly
higher heat of adsorption for Chimassorb 944 relative
to Irganox 1010, it was expected that the former would
eventually displace the Irganox 1010. This was shown to
be true by DRIFTS analysis of the silica sample from
the FMC cell which showed total coverage by Chimassorb 944. The heat change associated with this displacement was not readily detected due to there only
being a small dierence in the net heat change and the
slow rate at which the displacement progressed. The

Fig. 7. Raw FMC data showing adsorption of Irganox 1010 followed


by switch over to Chimassorb 944 (at 106 min): (a) thermal signal, (b)
dierential refractometer signal.

Fig. 6. DRIFTS spectra showing OH stretching region of silicas treated at 95 C in vacuo for 16 h: (a) Silica A, (b) silica B.

Fig. 8. Raw FMC data showing adsorption of Chimassorb 944 followed by switch over to Irganox 1010 (at 120 min): (a) thermal signal,
(b) dierential refractometer signal.

C.M. Liauw et al./Polymer Degradation and Stability 63 (1999) 391397

397

4. Conclusions

Acknowledgements

This study has highlighted the signicance of the


nature of ller surface and pore structure in prediction
of the adsorption of stabilisers, and any part that the
surface may play in potential controlled release of stabilisers into a polymer matrix.
The gel silica (Silica A) showed strong interaction,
from heptane, with polymeric HALS (i.e. Chimassorb
944) and large, multifunctional phenolic antioxidants
(i.e. Irganox 1010). This observation illustrated the signicance of multi-point adsorption. DRIFTS analysis
combined with FMC indicated that the strongest interactions with the silica surface were associated with,
amine and ester functionalities and to a lesser extent
phenolic OH groups. The precipitated silica (Silica B)
showed a far lower level of interaction with all the stabilisers, however, the Chimassorb 944 and Irganox 1010
still showed the strongest interaction.
Studies involving the adsorption of model compounds and examination of the silanol types/water content of both silicas by DRIFTS, indicated that Silica B
has a larger amount of hydrogen bonded water and
hence a larger number of Brnsted acid sites and vicinal
silanols than Silica A. This explained the greater activity
of Silica B towards amines and phenol and lower reactivity with ketones and esters. Silica A with its larger
proportion of Lewis acidic isolated silanols showed far
greater interaction with the latter two species, which
combined with its higher pore volume led to the greatest
overall adsorption activity.
Competitive adsorption studies of Chimassorb 944
and Irganox 1010 on to Silica A indicate the the former
will displace the latter from the silica surface in heptane.
If heptane is assumed to be a model for a polyolen
matrix, it is likely that Chimassorb 944 will sacrically
adsorb onto Silica A. Such an eect has been shown to
have important implications for stabilisation performance
under photo and thermo-oxidation. These ndings will be
reported separately.

The authors would like to thank Unilever Research,


Croseld Limited and the EPSRC for providing funding
for A. Childs.
References
[1] Hancock M. In: Rothon RN, editor. Particulate-lled polymer
composites. London: Longman Scientic, 1996. p. 282.
[2] Simmons DN. In: Blow CM, editor. Rubber technology and
manufacture. London: Butterworths, 1971. p. 192.
[3] Wolfschewenger J, Hauer A, Gahleitner MG, Neil W. Proceedings Eurollers 97, British Plastics Federation. Manchester, UK,
1997. p. 375.
[4] Childs A, Allen NS, Liauw CM, Franklin KR. Proceedings
Eurollers 97, British Plastics Federation. Manchester, UK,
1997. p. 371.
[5] Allen NS, Edge M, Corrales T, Childs A, Liauw C, Catalina F,
Pienado C, Minihan A. Polym Degrad Stab. 56 1997;125.
[6] Ashton DP, Rothon RN. In: Ishida H, editor. Controlled interphases in polymeric materials. New York: Elsevier Science, 1990.
p. 295.
[7] Liauw CM, Childs A, Edge M, Allen NS, Corrales T, Minihan
AR, Franklin KR. Proceedings Loughborough Fillers Symposium II. Loughborough University, UK, 1996.
[8] Liauw CM, Rothon RN, Hurst SJ, Lees GC. Composite Interfaces, in press.
[9] Jednacak-Biscan J, Cukman D. Colloids and Surfaces 1989;41:
87.
[10] Snyder LR. Principles of adsorption chromatography. New
York: Dekker, 1968. p. 156.
[11] Ross RA, Taylor AH. Proc Brit Ceram Soc 1965;5:167.
[12] Dzhgit OM, Kiselev AV, Muttik GG. Kolloidn Zh 1961;23:
504.
[13] Mauss M, Englehardt HJ. Chromatography 1986;371:235.
[14] Zhdanov SP. Dokl Akad Nauk SSSR 1949;68:99.
[15] Hertl W, Hair ML. Nature 1969;223:1151.
[16] Prigogine M, Fripiat JJ. Bull Soc Roy Sci Liege 1974;710:
449.
[17] Iler RK. The chemistry of silica. New York: Wiley, 1979. p.
632.
[18] Ashton DP, Briggs B. In: Rothon RN, editor. Particulate-lled
polymer composites. London: Longman Scientic, 1996. p. 96.
[19] Lange KR. J Colloid Sci 1965;20:231.

Potrebbero piacerti anche