Sei sulla pagina 1di 11

Why Silicones Behave Funny

M. J. Owen
Scientist Emeritus
Dow Corning Corporation

Abstract
This article provides the curious scientific reader with an overview of the surface properties of silicones, and
polydimethylsiloxanes in particular. From where do these surface properties originate, and why do they make silicones
useful in so many applications? This article is based on a previous publication from 1981 and includes comments
on progress made since at understanding these exceptional and funny surface properties.

Introduction
Since their introduction in 1943, silicones have been applied in an exceptional variety of industries. In many cases,
silicones have gained acceptance and are used commercially either as coatings or as additives because of their unique
surface properties. The focus of this article is on these specific surface properties and related applications, and it
concentrates on the physical property related aspects and their underlying molecular explanation. Hopefully it will
stimulate readers to consider how this unique property profile might be used in their specific area. This article is
based on a still-relevant 1981 article, that sought to answer the question Why are silicones used in so many surface
property linked applications? (see Note 1). This updated version is reprinted/republished with permission from
CHEMTECH, Copyright 1981 American Chemical Society [1].
Interested readers may also want to consider a recent book chapter that covers the topic in more detail [2].

Silicones in Surface Property Applications


As indicated above, silicones are used in many applications because of their surface properties, either as coatings
or additives (Table 1).
Table 1. Established Surface-Related Applications of Silicones
Applied coatings
Paper release
Water repellant
Electrical insulating compound
Pigment treatment
Fiber lubricant
Textile handle modifier
Mold release agent
Metal protector
Antifouling material
Masonry treatment
Anticaking aid

Integral additives
Polyurethane and other foam stabilizers
Integral plastics lubricant
Emulsifier
De-emulsifier
Wetting agent
Antifoam
Cosmetic formulation component
Polishing oil
Gloss enhancer
Powder coating flow-out aid
Paint additive

For coatings, the physical nature of the coating is important and cross-linking of silicone polymer chains is often
involved. As additives, diffusion to the interface is critical and so compatibility with the system is essential, often
requiring the use of silicone copolymers. Both the cross-linking and copolymerization aspects utilize the unique
chemistry of silicones, but it is not these synthetic aspects which chiefly account for the range of surface active
applications of silicones (see Note 2). Rather, it is a consequence of the unique physical properties of the silicones,
in particular polydimethylsiloxanes (PDMS), as I will focus on hereafter.
The best known of the commercially available silicones are polydimethylsiloxanes (PDMS), so this will be our main
concern. They are available as linear fluids, cyclics, gels and resins depending on the degree of cross-linking, or as
elastomers when fillers are incorporated in cross-linked polymers.
Table 2 lists both surface and bulk phenomena that together make up the unique property profile of PDMS. Structurally,
PDMS consists of an inorganic siloxane backbone with pendant methyl groups (Figure 1).

Table 2. Properties of Polydimethylsiloxanes


Low surface tension
Moderate water interfacial tension
No surface viscosity
Spreading and creep behavior
Variety of configurations
Large free volume
Low glass transition temperature
Low activation energy of viscous flow
Liquid nature at high molecular weight (linear polymers)
Small temperature variation of physical constants
Low boiling points (oligomers)
Low freezing and pour points
High compressibility
High permeability to gas and low molecular weight species
Low flammability and fire hazard
Low environmental hazard
Excellent weather resistance

Figure 1. A three-dimensional view of Me3SiO-(SiMe2O)4-SiMe3 showing the shielding of the polar main chain
made of Si-O groups by the methyl groups (structural representation courtesy of S. Grigoras, Dow Corning).

Inorganic silicate-like structures are associated with high surface energies, whereas hydrocarbon materials generally
have low surface energy. One idealized, but nevertheless useful, viewpoint is that the principal role of the backbone
in the surface activity of silicones is its ability to present the attached organic groups at interfaces. From this
perspective, the key property of the backbone is its flexibility and the key property of the organic group is its intrinsic
surface activity, or perhaps more fundamentally, the strength of the intermolecular interactions between these groups.
Polydimethylsiloxanes then emerge as a favored case of a very surface-active (very low intermolecular forces)
pendant group, methyl, whose activity is presented to best effect by virtue of the unique flexibility of the backbone
(Figure 2).

Figure 2. A water droplet on a cotton substrate coated with a polydimethylsiloxane, showing the overwhelming
importance of the methyl groups leading to hydrophobicity of the coated substrate (photograph courtesy
of A. Goodwin, D. Futter, A. Hynes and S. Leadley, Dow Corning Plasma Solutions).

In addition, the particular combination of siloxane and methyl has thermal and oxidative stability benefits. PDMS
is also essentially non-irritating and non-toxic. These features are exploited in several of the surface-active applications.
Let us now look further into the unique molecular architecture that is responsible for this behavior.
Flexibility of the silicone backbone. The flexibility of the siloxane backbone chain is unique. This assertion seemed
clear in 1981, but it needs more elaboration today, as newer, flexible perfluoroethers have become better known.
We can equate chain flexibility to freedom of rotation about bonds. The energy required for rotation about the carbonto-carbon bonds in polyethylene is 13.8 kJ/mol (3.3 kcal/mol). In polytetrafluoroethylene, this energy is greater than
19.6 kJ/mol (4.7 kcal/mol). In PDMS, this energy is almost zero, rotation being virtually free [3]. This freedom to
rotate is reflected in the glass transition temperature (Tg). It is not only the internal mobility in a polymer that
determines Tg; polymer free-volume, attractive forces between molecules, chain stiffness, and chain length all
contribute. Nevertheless, a low Tg indicates polymer flexibility; some comparative values are given in Table 3 [4].
Table 3. Selected Values of Glass Transition Temperature

Poly(pentamethylcyclopentasiloxane) (PD5)
Co-poly(oxytetrafluoroethylene-oxydifluoromethylene)
Polydimethylsiloxane
Polyethylene
Poly(thiodifluoromethylene)
Cis poly(1-pentenylene)
Poly(oxypropylene)
Polyisobutylene
Polymethyltrifluoropropylsiloxane
Polytetrafluoroethylene

Tg (K)
122
140
146
148
155
159
198
200
203
293

I have added two values to Table 3; that for co-poly(oxytetrafluoroethylene-oxydifluoromethylene) or co-poly(CF2CF2OCF2O) [5] and poly(pentamethylcyclopentasiloxane) or PD5 [6]. It is no longer true that PDMS has the lowest known
Tg; clearly this perfluoroether and PD5 are lower. This latter silicone polymer is a methylsiloxane composed of linked
rings. It is formed from (MeHSiO)5 or D5H, a cyclic siloxane pentamer with a methyl group and hydrogen atom on
each silicon. All polymers with a Tg less than 160 K, except the silicones PDMS and PD5, are pure backbone
polymers with no pendant groups. Once pendant groups are introduced, Tg rises considerably. Evidently, the siloxane
backbone is still unusually flexible. In the original CHEMTECH article I cited Langmuir trough studies with
transitions between different spread configurations occurring with a very small difference in surface pressure. I think
this is still good evidence of chain flexibility although there has been considerable rethinking in the past two decades
concerning the nature of spread PDMS surface layers. A fuller account of these developments is given in reference 2.
Table 4 lists fundamental polymer architectural geometry parameters, bond angles and bond lengths, taken from
model compounds [7]. These angles and lengths are from hexamethyldisiloxane and the two organic systems that
PDMS is most often compared to, polyethers and hydrocarbons.
Table 4. Model Compound Chain Parameters

Hexamethyldisiloxane
Dimethylether
Propane

Bond
Si-O
C-O
C-C

Length (nm)
0.163
0.142
0.154

Bond
Si-O-Si
C-O-C
C-C-C

Angle ()
130
111
112

The siloxane system has the most open structure with the flattest angle and the longest bond length. In most
hydrocarbon systems, bond angles are fixed, i.e., tetrahedral. Thus, available surface-active groups, such as methyl,
cannot adopt the lowest surface-energy configuration. For example, in polyisobutylene, despite an exceptionally
large backbone bond angle of 123, only a helical structure is possible [8]. In contrast, the Si-O-Si angle shows much
4

wider variability. This contributes to the variety and ease of inter-conversion of configurations. Values of 105-180
have been quoted [9]. Widening of the angle when silicon replaces carbon was formerly attributed to (p d) bonding
[10] but is now known to reflect a strong deviation from sp3 hybridization [11, 12]. The unusually wide bond angle
of the Si-O-Si bond reflects a strong deviation from sp3 hybridization. This is due to the transfer of electronic charge
from the oxygens lone pairs to the bonding zone between silicon-oxygen atoms. This charge transfer results in a
shorter Si-O bond length than expected, and the alteration of the value of the bond angle at oxygen as a consequence
of the attenuation of the sp3 hybridization. These conclusions were drawn by subtracting the electron charge density
calculated at ab initio level with two basis sets: one that estimates correctly the geometry of this moiety, and one
that predicts a standard sp3 hybridization. Furthermore, the ab initio results indicate that the equilibrium value for
the Si-O-Si bond angle is around 145-160, and the energy required to deform this bond angle to 180 is very low,
approx. 1.3 kJ/mol (approx. 0.3 kcal/mol). However, the energy to deform this bond angle to lower values, towards
109 (sp3 hybridization value) increases very quickly for low angle-variations.
Another structural factor that must enhance chain flexibility is the unsubstituted alternating oxygen linkage that
much reduces steric interference possibilities during reorientation. PDMS is still by far the lowest Tg polymer of
this [Me2XY]n structure [2].
Methyl group surface energy. Here we will assume that there is an intrinsic surface energy for a methyl group,
regardless of whether it is pendant to a hydrocarbon or siloxane chain (i.e., intermolecular forces between methyls
are essentially the same in both cases). Our assumption is based on the similarity in critical surface tension of wetting
of paraffin wax (23 mN/m) [13] and PDMS (22.7 mN/m) [14]. Both these situations are believed to produce arrays
of closely packed methyl groups at the surface. It is a well-known result of Zismans studies [15] that the order of
increasing surface energy for single-carbon based moieties is:

CF3- < -CF2- < CH3- < -CH2Much of the relative surface energy behavior can be understood with this order. It explains why perfluoroethers,
such as poly(perfluoropropylene oxide), with many pendant CF3- groups, have a lower surface energy than the
perfluoroalkanes. It also explains why polydimethylsiloxanes have a lower surface energy than the alkanes.
Pauling quotes the van der Waals area of a methyl group as 0.12 nm2 [16]. Zisman gives 0.227 nm2 for a dimethylsiloxane
unit in a spread configuration [17]. This suggests that two methyl groups fit closely over each siloxane linkage. The
CF3- group might give a more surface-active siloxane polymer but, being larger, might not be able to align for maximum
efficacy. In any case, a hydrocarbon bridge, e.g., a (-CH2-CH2)- group, has to be placed between the CF3- and the
siloxane backbone to achieve adequate chemical stability. These -CH2- groups offset the CF3- groups effect to the
extent that polymethyltrifluoropropylsiloxanes have higher liquid surface tensions than PDMS. This may also be partly
due to the quadrupole (uncompensated dipole) at the CF3-CH2- hydrocarbon/fluorocarbon junction.
Other important aspects. At 22.7 mN/m, the critical surface tension of wetting of PDMS is higher than its liquid
surface tension, which is 20.4 at 20C at the highest molecular weight measured [18]. Consequently, the polymer
will spread over its own adsorbed film. This creep of silicone fluid can be a problem if unreacted fluid is present
in silicone compositions and migrates to electrical contacts, as PDMS is a good electrical insulator. On the other
hand, it can be an advantage in achieving complete surface coverage in applications such as metal protection,
pigment surface treatments, insulator compounds and mold release. Low surface energy is responsible for many
of the applications of PDMS. For example, it serves as the hydrophobic entity in surfactants that need to reduce
the surface tension of water to a lower level than is possible with conventional surfactants [19].
In many cosmetics, silicones provide emolliency, not so much for their lubricity, but for their outstanding spreadability
[20]. In medical devices, they are used to siliconize needles and reduce pain when penetrating the skin, again because
of their spreadability, and adequate lubricating properties, not to mention their biocompatibility [21].
One fascinating aspect of the particular surface energy of PDMS is that the most biocompatible range of polymers
seems to be at surface energies of 20-30 mN/m [22]. The original use of silicones in biomedical applications has
much to do with their position in this range (see Note 3). It is postulated that this range has a minimum interfacial
tension with protein-containing aqueous solutions. This results in the smallest possible driving force for protein
adsorption and denaturation at these values (Figure 3).
5

Figure 3. Comparison of a medical grade polycarbonate without (left) or with (right) a silicone coating,
both after exposure to human blood for 5 minutes, washing and electron microscopy analysis and showing
the reduced amount of cell and protein absorption achieved by the silicone coating (pictures courtesy of
F. Briquet and C. Micquel).

Note that it is not materials, which are totally wetted by water that are usually particularly biocompatible. This point
implies the significance of nonpolar dispersion force interactions in this instance. These considerations are also
important in other fields, such as marine antifouling.
There are two further consequences of the low intermolecular forces that result from the molecular architecture of
polydimethylsiloxane. These are its low surface shear viscosity and high gas permeability. PDMS has the lowest recorded
surface shear viscosity [23] and its oxygen permeability, for example, is second only to another class of organosilicon
polymers, the trimethylsilyl alkynes such as poly(1-trimethylsilyl-1-propyne) [24]. There are applications where both
high gas permeability and low surface viscosity are exploited, for example as antifoams in fermentation processes.
It is not possible to cover all the applications for PDMS that depend, at least in part, on this unique combination
of surface properties. By now, interested readers should be able to see how silicones might fit into their interests.
Instead of any attempt to be comprehensive, lets confine ourselves to four major, and apparently very contradictory,
applications to illustrate the versatility of these polymers: antifoams, polyurethane foam stabilization, release coatings
and pressure-sensitive adhesives.

Applications
Antifoams. The control or elimination of foam in many industrial processes is a major application for PDMS fluids.
Typical examples are in textile dyeing and finishing, in gas scrubbing at petrochemical plants, in penicillin production
by fermentation and in the formulation of controlled foaming detergents.
Foaming originates from the presence of a surfactant, which stabilizes the foam films by a variety of mechanisms.
Chief among these are surface elasticity and surface viscosity. Materials that diminish or eliminate these mechanisms
are known as antifoams or defoamers. PDMS provides low surface and interfacial tensions, thus achieving the
prerequisite condition of being more surface active than the foaming surfactant they are destined to replace. They
do this by entering and spreading in the foam films, without conferring any additional direct foam-stabilizing
mechanisms such as high surface viscosity. These properties are sufficient to enable silicones to be effective
defoamers for nonaqueous systems. To be active at low concentrations, the antifoam should be insoluble in the
foaming medium, although this is not an absolute necessity. Other ancillary properties can also be important: for
example, high gas permeability and lack of toxicity are vital in fermentation processes.
For aqueous systems, it has long been recognized that finely dispersed hydrophobic solids, such as polydimethylsiloxane-treated silica, are needed for effective antifoam action. Such fluid/solid combinations are known as
antifoam compounds (Figure 4).

Figure 4. Because silicone antifoams are effective at very low levels (often 10-100 ppm), their use can
eliminate the need for much greater quantities of hydrocarbon antifoams. Excessive use of hydrocarbon
antifoams can produce problems of product contamination or excessive biological oxygen demand in
industrial effluents. Note the small red circle around the silicone antifoam drop, the middle blue circle where
defoaming has occurred and the larger green circle where spreading has occurred but not yet defoaming
(photograph courtesy of J.P. Lecomte, Dow Corning).

They are often prepared as emulsions or encapsulated granules for ease of application. It is now known that these
hydrophobic solids play an important role in the breakdown of the so-called pseudo-emulsion film that forms between
the antifoam droplet and the surface of the foaming solution [25]. Bridging effects across the two surfaces of a foam
film leading to rupture are also significant.
Polyurethane foam stabilization. A prerequisite for foam stability is positive adsorption of a surface-active species.
In polyurethane foam formation, the bulk of the liquid mixture is often a polyol, such as the poly(propylene oxide)
adduct of glycerol, which has a surface tension of approximately 32 mN/m. This surface tension is the lowest many
organic surfactants can achieve and, consequently, they are not useful as stabilizers for polyurethane foam.
Polydimethylsiloxanes, with their potential for lowering surface tension a further 10 mN/m, are ideal surfactants
for such systems. They are made compatible with the polyurethane foam mix by copolymerizing silicones with
poly(alkylene oxides) (Figure 5).

Figure 5. Without the controlled thinning of cell walls made possible with silicone surfactants, high-quality
polyurethane foams would not be possible: left without a silicone poly(alkylene oxide) copolymer surfactant,
right with one such surfactant and showing the improved cell size quality (pictures courtesy of M. Stark ,
M. Ferritto and M. Stanga, Dow Corning).

There is now a wide range of such copolymers. Their structures and compositions are tailored to give optimum
foam-supporting film viscoelasticities for todays wide variety of polyurethane foam systems [26]. This type of
7

copolymer is also used in other polymer systems, e.g., phenolic foams. In some high temperature applications, such
as textile dyeing where the cloud point is exceeded so that these surfactant copolymers become insoluble, they also
function as antifoams.
Poly(alkylene oxides) are not the only water-solubilizing entities that can be attached to a silicone hydrophobe.
Indeed, silicone analogs of most conventional hydrocarbon surfactants are possible. By varying the HLB (hydrophilelipophile balance), a range of useful silicone surfactants is possible including powerful wetting agents for low energy
surfaces and water-in-oil emulsifiers.
Release coatings. Silicones are widely used as release coatings. The low surface energy and low cohesive strength
conferred by the methyl groups combine to provide the excellent release characteristics of PDMS fluids. They spread
rapidly on metal mold surfaces and have excellent thermal stability. This combination of properties allows them to
be used at rubber-curing temperatures, thus providing an efficient parting agent between organic polymer and mold
surface. A blend of high viscosity silicone fluid with a lubricating particulate solid such as mica provides the difficult
combination of slip, air-bleed, release and thermal stability demanded by the tire industry.
Perhaps the most familiar release application for silicones is as release liners for pressure-sensitive adhesive coated
labels and the like. In this application, it is important that no gross transfer of silicone occurs to interfere with the
subsequent adhesion of the pressure-sensitive adhesive. Thus, cross-linked PDMS films are used (Figure 6).

Figure 6. The low surface energy of silicone paper coatings on the release sheet allows the do-it-yourselfer
to easily handle aggressive pressure-sensitive adhesives such as those used on these labels (photograph
courtesy of Luc Dussart, Dow Corning).

For good release, it is important to have a lower surface energy substrate than the adhesive. Typical acrylic and SBR
adhesives have surface tensions of 30-40 mN/m, well above that of polydimethylsiloxanes. The low substrate surface
energy not only directly lowers the thermodynamic work of adhesion, but also produces poor wetting, i.e., a high
contact-angle between adhesive and substrate, which not only reduces interfacial area but also concentrates the
stresses in the subsequent separation of the adhesive/release-coating laminate.
One of the key new insights in the release area is the discovery that slippage can occur between the adhesive layer
and the silicone release layer [27]. Such behavior has ramifications in other applications. One important example
is in the significant use of silicones in creams and lotions in the personal care area. The benefits of a low surface
energy, water insoluble polymeric component are obvious. The silicone will spread easily, repel aqueous solutions
and remain on the surface of the skin where topically applied with no tendency for absorption. However, a key
attraction to the customer of the silicone is a much less quantifiable attribute, the so-called smooth, soft or silky
touch of the treated skin. A similar phenomenon is experienced with silicone hair treatments and the soft handle
of silicone-treated textiles. Facile slippage of finger surfaces over silicone-treated skin is likely another manifestation
8

of this tendency to interfacial slippage or low static coefficient of friction. It seems to be a consequence of both the
fundamental attributes stressed in this article: low intermolecular forces between methyl groups and the very high
flexibility of the siloxane backbone.
Pressure-sensitive adhesives. The first requirement of adhesion is good wetting. This will clearly be a significant
positive factor with such a low-surface-energy adhesive. However, an ability to spread on a wide variety of substrates
is only one part of the requirement; pressure-sensitive tack and cohesive strength are also needed. I will not go into
the details of these bulk properties, but let me just state that a balance between these requirements must be achieved.
This is done by formulating reactive fluid-resin mixtures where the PDMS fluid and the tackifying resin contain
groups that will react by silanol condensation to form a network of adequate strength [28].
Silicone pressure-sensitive adhesives have a number of advantages over conventional organic adhesives. They are
chemically and thermally stable and are resistant to moisture and weathering. They adhere well to many types of
surfaces such as metals, glass, paper, fabric, plastics and human skin, and to materials such as silicone elastomers and
release coatings that are usually expected to be nonadhesive. A typical example is silicone pressure-sensitive adhesive
used in drug-loaded transdermal patches and associating many of the silicone characteristics mentioned so far: adhesion
to an hydrophobic substrate (here the skin), allowing permeability for the drug to diffuse through the adhesive into
the skin because of the low Me-to-Me intermolecular interactions, not to mention the lack of skin irritation associated
with silicones that makes these adhesives particularly suited when long wear time is considered (Figure 7).

Figure 7. Silicone pressure-sensitive adhesives are used in transdermal drug delivery systems to fix medicated
patches on the skin, because these adhesives allow drug permeation and do not create skin irritation
(photograph courtesy S. Postiaux, Dow Corning).

It also means that silicone pressure-sensitive adhesives cannot be conveniently packaged and delivered with
conventional release coatings. Even lower surface energy coatings based on fluorosilicones have had to be developed
to accomplish this [29].

The Future
What does the future hold for silicones in this area? The future of 1981 is todays history, so the value of my
predictions is readily apparent. In the original article I safely predicted a continuous broadening of their use in
surface-active applications based on the unique combination of properties I have described. I am happy to report
that this prediction has come true. Table 5, taken mostly from reference 2, lists examples of applications that were
well on the horizon in 1981, if not unrecognized. We anticipated that the low toxicity of PDMS would be a considerable
factor in this growth, as would the fact that polydimethylsiloxanes do not accumulate in the environment and
ultimately break down to innocuous natural products such as carbon dioxide and silicic acid [30]. The coatings and
foam applications listed in Table 5 are evidence of the validity of this assertion.
9

Table 5. New and Emerging Surface-Related Applications of Silicones


Poly(amidoamine-organosilicon) (PAMAMOS) dendritic coatings
Silsesquioxane precursors for ceramic coatings
Tailored, tethered surface-modifying additives
Nontoxic, biofouling release coatings
Microemulsions
Organosilicon gelling agents
Microcontact printing with PDMS elastomers
Electrically conductive silicone adhesives
Controlled-foaming, environmentally-friendly, low-temperature detergents
Wound dressings
Polysiloxane is the most flexible chain available, but the aliphatic fluorocarbons provide a more surface-active
pendant group than methyl. There was already considerable literature on fluorosilicones in 1981; consequently, an
easy accurate prediction was the commercialization of silicones containing more fluorine than those available two
decades ago. The use of polymethylnonafluorohexylsiloxane as a release coating for PDMS-based pressure-sensitive
adhesives is the prime example of this ongoing trend [29].
Many developments were anticipated in the copolymer area. We have discussed only copolymerization with
poly(alkylene oxides). But there were, and continue to be, numerous reports of other silicone-organic block copolymers
that should lead to future commercial products, such as additives to modify surface properties, or be useful new
materials in their own right. There has been much interest in this topic in recent years but it is fair to say that this
has been mostly of a scientific nature and the full commercial promise of this prediction has yet to materialize.
Similarly we predicted the use of PDMS to enhance recovery of finite fossil fuels, which has materialized in the
form of de-emulsifiers on off-shore extraction rigs, and silicone/natural product hybrids, which still remain tantalizingly
just in the future.
The resistance of the silicone surface to weathering has proved to be a key factor in a future more concerned with
conservation than with built-in obsolescence. We envisaged new construction components, coatings and additives
in solar energy devices but missed the big development of silicone high-voltage insulators, and who would have
predicted a large current market in air bags based on reliability of deployment attributes two decades ago?
We can, however, still stand by our original conclusion. The diversity of silicone applications is already remarkable,
but it still has a long way to go before it is finished. It has come about because of the unusual combination of
properties and attributes I have discussed. These attributes guarantee an even more astonishing future. In six decades,
silicones have become indispensable as auxiliaries in many different branches of industry. There is every reason to
expect the next 60 years of silicones to be even more exciting than the past.

Notes:
1. I have tried to change only what is absolutely necessary in my original article and based on over two decades
of progress in this field. In particular, I have retained the original title. This was assigned by the CHEMTECH editor
without consultation. It offended me at the time as my British eyes required the adverbial form funnily. Now, after
much longer in the USA, I have grown used to it.
2. In the interests of space I have omitted a number of surface-related facets of organosilicon chemistry, of which
the most important is perhaps the silane coupling agents. The interfacial chemistry of these materials is complex
and fascinating [31], but it falls outside the basic structural theme considered here.
3. The original article used success in biomedical applications rather than original use. I had not anticipated
the breast implant debacle. I remain convinced that these concerns are not supported by the relevant science and am
optimistic that science will eventually prevail over the fear, bias and greed that has recently driven much of this debate.
10

Acknowledgments
My thanks to Laurence Bartier for computerizing my original
article, to Dr. Stelian Grigoras for his comments about the siloxane
electronic hybridization, and to Dr. Andr Colas for suggesting the
updating of my original article and for his insightful editorial
comments (all from Dow Corning).
This article was published in Chimie Nouvelle, 85, 27 (2004), and
is reproduced here with the kind permission of the editor.

References
1.
2.

Owen, M. J., CHEMTECH, 11, 288 (1981).


Owen, M. J., in Silicon-containing polymers, Jones, R. G.,
Ando, W. and Chojnowski, J., Eds.; Kluwer Academic
Publishers, Dordrecht, 213 (2000).
3. Tobolsky, A. V., in Properties and structures of polymers,
John Wiley and Sons, New York, 67 (1960).
4. Lee W. A. and Rutherford, R.A., in Polymer handbook 2nd
ed., Brandrup, J. and Immergent, E. H., Eds., John Wiley and
sons, New York, III-139 (1975
5. Scheirs, J., in Modern fluoropolymers, Schiers, J. Ed., John
Wiley and Sons, New York, 435 (1997).
6. Kurian, P., Kennedy, J. P., Kisliuk, A. and Sokolv, A., J.,
Polym. Sci. Part A, Polym. Chem. Ed., 40, 1285 (2002).
7. Tables of interatomic distances and configurations in
molecules and ions, Spec. Publ. No. 11, The Chemical Society,
London (1958).
8. Benedetti, E., Pedone, C. and Allegra, G., Macromolecules,
3, 16 (1970).
9. Voronkov, M. G., Mileshkevich, V. P. and Yuzhelevskii, Y. A.,
in The siloxane bond, Livak J., Transl., Consultants Bureau:
New York, 12 (1978).
10. Ebsworth, E.A.V., in "Organometallic compounds of the group
IV elements"; MacDiarmid, A.G. Ed., Marcel Dekker, New
York, 1 (part 1), 1 (1968).
11. Grigoras, S., Lane T.H. In Conformational analysis of
substituted polysiloxane polymers in Silicon-based polymer
science, Zeigler, J.M. and Fearon, G.F.W. Eds, American
Chemical Society (ACS Adv. Chem. Ser. 224), Washington
DC, Chap 7, 125 (1990).
12. See comments in Brook, M.A. Silicon in organic,
organometallic, and polymer chemistry, John Wiley & Sons,
Inc, 33 (2000).

13. Owens, D. K. and Wendt, R. C., J. Appl. Polym. Sci., 13, 1741
(1969).
14. She, H., Chaudhury, M. K. and Owen, M. J., in Silicones and
silicone-modified materials, Clarson, S. J., Fitzgerald, J. J.,
Owen, M. J. and Smith, S. D., Eds.; ACS symposium series
No. 729, 241 (2000).
15. Zisman, W. A., in Advances in chemistry series, no. 43,
American Chemical Society, Washington DC, 1 (1964).
16. Pauling, L., in The nature of the chemical bond, 3rd ed.,
Cornell U.P., 257 (1960).
17. Fox, H. W., Taylor P. W. and Zisman, W. A., Ind. Eng. Chem.,
39, 1401 (1947).
18. Roe, R. J., J. Phys. Chem., 70, 2013 (1968).
19. Hill, R. M., in Silicone surfactants, Hill, R. M. Ed., Surfactant
Science Series, 86, Marcel Dekker, New York, 1 (1999).
20. Brand H.M. and Brand-Garnys, E.E., Cosmetics & Toiletries,
107, 93 (1992).
21. Smith, E.J. (chairman), in Siliconization of parenteral drug
packaging components, Technical report no. 12, J. of
Parenteral Science and Technology, 42, S1 (1988).
22. Baier, R. E., in Adhesion in biological systems, Manly, R.
S. Ed., Academic Press, New York, 15 (1970).
23. Jarvis, N. L., J. Phys. Chem., 70, 3027 (1966).
24. Nagai, K. and Nakagawa, T., J. Membrane Sci., 105, 261
(1995).
25. Hill, R. M. and Fey, K. C., in Silicone surfactants, Hill, R.
M. , Ed., Surfactant Science Series, 86, Marcel Dekker, New
York, 159 (1999).
26. Snow, S. A. and Stevens, R. E., in Silicone surfactants, Hill,
R. M., Ed., Surfactant Science Series, 86, Marcel Dekker,
New York, 137 (1999).
27. Newby, B. Z., Chaudhury, M. K. and Brown, H. R., Science,
269, 1407 (1995).
28. Merrill, D. F., Adhesives Age, 22, 39 (1979).
29. Maxson, M. T., Norris, A. W. and Owen, M. J., in Modern
fluoropolymers, Schiers, J. Ed., John Wiley and Sons, New
York, 359 (1997).
30. Xu, S., Lehmann, R. G., Miller, J. R. and Chandra, G., Environ.
Sci. Technol., 329, 1199 (1998).
31. Owen, M. J., in Surfaces, chemistry and applications,
Chaudhury, M., Pocius, A. V., Eds., Adhesion Science and
Engineering - 2, Elsevier, New York, 403 (2002).

LIMITED WARRANTY INFORMATION - PLEASE READ CAREFULLY


The information contained herein is offered in good faith and is believed to be accurate. However, because
conditions and methods of use of our products are beyond our control, this information should not be used in
substitution for customer's tests to ensure that Dow Corning's products are safe, effective, and fully satisfactory
for the intended end use. Suggestions of use shall not be taken as inducements to infringe any patent.
Dow Corning's sole warranty is that the product will meet the Dow Corning sales specifications in effect at the
time of shipment. Your exclusive remedy for breach of such warranty is limited to refund of purchase price or
replacement of any product shown to be other than as warranted. DOW CORNING SPECIFICALLY DISCLAIMS
ANY OTHER EXPRESS OR IMPLIED WARRANTY OF FITNESS FOR A PARTICULAR PURPOSE OR
MERCHANTABILITY. DOW CORNING DISCLAIMS LIABILITY FOR ANY INCIDENTAL OR CONSEQUENTIAL DAMAGES.

We help you invent


the future.
www.dowcorning.com

Dow Corning is a registered trademark of Dow Corning Corporation.


WE HELP YOU INVENT THE FUTURE is a registered trademark of Dow Corning Corporation.
2005 Dow Corning Corporation. All rights reserved.
Printed in USA

VIS2420

Form No. 01-3078-01

Potrebbero piacerti anche