You are on page 1of 17

Chemical and physical modification of fluoropolymer

surfaces for adhesion enhancement: a review

LORRAINE M. SIPERKO1,* and RICHARD R. THOMAS2


1IBMSystemsTechnologyDivision, Advanced TechnologyLaboratory, Endicott, NY13760, USA
2IBMResearch Division, Thomas J. WatsonResearch Center, YorktownHeights, NY 10598,USA

Revised version received 18 November 1988

Abstract�The primary goal of surface modification of fluoropolymers is to improve adhesion to these


low surface energy materials. Both classical methods and newer processes have been used as a means
to this end. Various methods of modifyingfluoropolymer surfaces include wet chemical etching, elec-
trochemical reduction, grafting, application and removal of metals, ion and electron beam techniques,
and plasma modification. In addition to the modification procedure, it is necessary to evaluate the
effectiveness of the chosen method by subsequent analysis of the modified surface. Physical analytical
methods include contact angle and wettability measurements, lap shear and composite tensile shear
strengths, peel strengths, and surface topographical determinations. Chemical analyses used include
infrared, Raman, Rutherford backscattering, ultraviolet-visible, wide angle X-ray scattering, X-ray flu-
orescence, and X-ray photoelectron spectroscopies as well as thermal desorption mass spectrometry.
Each of the modification methods, with results of the subsequent chemical or physical analysis, will be
discussed.

Keywords:Fluoropolymer; surface modification;adhesion; bonding.

1. INTRODUCTION .
The science of adhesion is multi-disciplinary. Aspects which must be considered
are surface chemistry and physics, rheology, polymer chemistry and physics,
stress analysis, and fracture phenomena. It should not be surprising, then, that
the mechanism of adhesion is not fully understood, and various mechanisms have
been proposed [1].
If one considers thermodynamic criteria, it is easy to see that the 'best'
adhesion occurs when the interfacial free energy is at a minimum. This energy
will be increased by the presence of atomic-scale disorder and dislocations at the
interface. Reduction of this energy may be accomplished by relaxations such as
inter-atomic bonding across the interface [2]. These bonds may be physical or
chemical in nature and are often termed 'weak' or 'strong', respectively. The
terminology arises from the energy of the interatomic and intermolecular forces
[3].
The deleterious effects of surface contaminants and water on the interfacial
bond strength lead to the conclusion that chemical interactions at the interface
are important. It is known that contaminants of the order of 10-' g/cm2 can

*To whom correspondence should be addressed at: IBM T43/257-2A, 1701 North Street, Endicott,
NY 13760, USA.
158

adversely affect adhesion [4]. When considering polymer surfaces, several


characteristics are important [1, 5], and there is considerable emphasis on the
chemical and physical alteration of polymer surfaces in an effort to enhance
adhesion. These chemical and physical changes are the result of acid etching,
oxidation [6], ion implantation [7, 8], plasma [9] and surface abrasion [3].
Organic functional groups exist on the surface of chemically modified polymers
[6, 9]. Such surface groups may interact chemically with metals or other
polymers. The importance of metal-oxygen-polymer complexes has been
disputed [10, 11]; however, a correlation between bond strength and percent d
valence bond saturation for transition metals has been found [ 12].
In all instances, it should be noted that results of surface studies may not be
representative of the bulk and vice versa [ 13]. Attempts to measure adhesion are
repressed by two difficulties: (1) the inability to obtain perfect contact between
the substrate and the adhering film, and (2) application of a stress such that the
force required to separate the two materials may be accurately determined [10].
A recent article, however, addresses techniques for adhesion measurement of
films [14].
Fluoropolymers are known for their chemical inertness and thermal stability,
and are typically utilized in applications where these qualities are desired [9].
Their excellent insulating properties make them useful in electronic packaging
[ 15]. They also have use as coatings for spacecraft optics [ 16] and their ability to
discharge in simulated space environments is of interest [16-18]. One of the
unique properties of fluoropolymers is their low surface tension [ 19], resulting in
a low work of adhesion between most polar liquids and the fluoropolymer
surface. Fluoropolymers in general, and polytetrafluoroethylene in particular,
also have unique antifrictional properties [20] requiring pretreatment of the
surface to enhance adhesion.
Due to their low surface tension, the adhesion of high surface tension
materials, such as metals, to fluoropolymers is negligible. For many years, this
fact precluded the use of these types of materials for various engineering and
technological applications. Surface modification of fluoropolymers was thought
to be useless due to the inert structure of the polymer. In 1956, however,
researchers at the Diamond Ordnance Fuze Laboratories reported that the
surface of a Teflon'* coated stirrer, used in an organic reduction reaction with a
mixture of sodium and naphthalene in tetrahydrofuran, had been darkened [21],
indicating a reduction of the fluoropolymer surface to a graphite-like material.
Resulting bond strengths, using an epoxy adhesive in tension or shear mode,
ranged from 75 to 140 kg/cm2. Shortly after this initial finding, two patents
dealing with the chemical reduction of Teflon surfaces to increase adhesion were
granted [22, 23] and efforts to explain the increased adhesion of fluorocarbons
in contact with solutions of sodium in liquid ammonia or naphthalene followed.
The original sodium/naphthalene work showed that the surface of fluoro-
polymers may indeed be chemically treated with resultant changes in both the
chemical and the physical properties of the treated surface. Succeeding experi-
mentation has shown that a wealth of methods for altering fluoropolymer

*Teflon, Teflon FEP, PFA, and Tefzel are trademarks of E.I. duPont de Nemours and Co. of
Wilmington, DE.
159

surfaces exists, leaving the researcher to determine which is the most convenient
or effective method for a given process.
In the following discussion, the aspects of fluoropolymer adhesion enhance-
ment are grouped into four major categories: chemical modification, crystalliza-
tion and crosslinking by metals, ion and electron beam modification, and glow
discharge plasma modification. The categories are representative but by no
means complete. The reader is referred to the references cited for additional
information regarding the modification of fluoropolymer surfaces.

2. CHEMICAL MODIFICATION

X-Ray photoelectron spectroscopy (XPS) and contact angle measurements were


used to examine Teflon FEP® (tetrafluoroethylene/hexafluoropropylene co-
polymer) before and after various treatments [24]. From the data shown in Table
1, it can be seen that major changes occurred on the Teflon FEP surface. Contact
angle data indicate that the Na/NH3 etched surfaces are much more hydrophilic
than the untreated Teflon FEP. Removal of fluorine (indicated by absence of the
F 1 s peak) leaves unsaturated sites, most likely C=C or C= C, which react with
bromine. Oxygen present in the reduced layer (indicated by the presence of an 0
ls peak) is probably in the form of >C=O, confirmed by bands at 1600-1799
cm-' in the infrared attenuated total reflectance spectrum. This oxidized hydro-
carbon-type surface is apparently unstable since abrasion or exposure to heat
and light may remove some of the layer to expose the underlying fluoropolymer.
It has also been observed, by thermal desorption mass spectrometry (TDS), that
low molecular weight fluorocarbons are desorbed as a sample of similarly treated
Teflon is heated [25]. The Tetra-Etch® treatment* is effective in removing lower
molecular weight fluorocarbons as evidenced by elimination of the mass 69
fragment during TDS measurements.
Other chemical methods have been successfully used to modify fluoropolymer
surfaces. Benzoin dianion, K2{PhC(0)C(0)Ph}, produces a surface with a
metallic gold color which fades after 1-2 days in air [26]. The material has a
wavelength maximum at 540 nm in its UV/visible spectrum, and loss of - 3.8
fluorine atoms per monomer unit (determined by gravimetric analysis) was
-
Table 1.
XPS and water contact angle data after various PTFE surface treatments [24]

°A 5.25% aqueous sodium hypochlorite solution.


bBr 3d signalat 67 eV detected.

*Tetra-Etch is the tradename of a Teflon surface treatment manufactured by W. L. Gore and


Assoc., Newark, DE.
160

confirmed by XPS. These samples react readily with C'2 or Br2, as confirmed by
infrared spectroscopy, and degradation of the surface was indicated by scanning
electron microscopy (SEM). Polyacetylene formation from the benzoin dianion
reduction of Teflon has been confirmed by Raman spectroscopic studies [28].
Teflon reduced in DMSO (dimethyl sulfoxide) or DMSO-d6 extracts solvent
hydrogen and deuterium, respectively. Raman spectra recorded continuously
over a period of 3 h in air showed reduction of the bands assigned to trans-
polyacetylene while bands associated with virgin Teflon appeared. The presence
of bands at 1600 and 2150 cm-' in photoacoustic infrared spectra indicate
aromatic rings and C=C, respectively. Similar spectroscopic studies with
PCTFE (polychloro-trifluoroethylene) show no evidence for trans-poly-
acetylene, but do indicate C= C bands. Hydrogen containing fluoropolymers can
be reduced by base-promoted dehydrohalogenation into polymers having
polyene segments [28]. The benzoin dianion reduced Teflon can be donor-doped
with Na or K naphthalenide in an inert atmosphere to give a surface with a
conductivity of 4.8 x 10 - 2 S2/cm, a much lower value than donor-doped trans-
polyacetylene. This could be due to a reaction of the dopant with the reduced
surface or to the small (12-28 units) conjugation lengths in the trans-(CH),. x
component.
Non-traditional reducing agents such as methyllithium, used on PCTFE, result
in reduction of the polymer with concurrent side reactions [29]. Contact angle
and mass loss data, shown in Table 2, suggest that the surface has become more
hydrophilic. This is confirmed by XPS results, which indicate that oxygen has
been incorporated into the polymer while chlorine and fluorine have been
depleted. There was contact angle hysteresis but surface roughening could not be
confirmed by SEM. Chemical reduction of the surface is demonstrated by the
appearance of an absorption band at 270 nm, indicative of conjugated double
bonds formed by the reductive removal of fluorine. Infrared studies of the
reaction of methyllithium with a PCTFE oil (lower molecular weight analog of
film) indicate that chemical grafting of methyl groups has occurred and a-
fluorocarbonyl groups and conjugated C=C bonds have been introduced into the
backbone structure of the oil. With a series of organolithium or Grignard
reagents, organic functional groups, such as carboxylic acid, aldehyde, and
alcohol functional groups, can be grafted onto the Teflon surface [30], which
should drastically alter their wetting and adhesion properties. Hydroboration
followed by oxidation produces hydroxyl groups, and bromination followed by

Table 2.
Chemical and physical effects of methyllithiumreaction time on PCTFE 1291

°Using water.
bAbsorbance at 270 nm.
161

liquid ammonia introduces amine groups onto a benzoin dianion-reduced Teflon


surface [31]. An interesting observation was made when sodium salts of
arylsulfides were allowed to react with saturated fluorocarbons. The aryl sulfides
not only replaced the fluorines but also introduced conjugation into the
backbone [32].
Alkali metal amalgams have also been successfully used to modify fluoro-
polymer surfaces. Lithium amalgam at 100°C produces a dull black, brittle
material which exhibits the properties of ceramics [33]. The reaction product has
good electrical conductivity and a graphitic-like structure [34]. It is thought that
carbon radicals react with Li+ to form CxLi compounds which are hydrolyzed on
contact with water to produce LiOH and C-H compounds. The films are
microporous solid adsorbents [35], and exhibit typical adsorption isotherms.
The measured surface areas are large (3300-4150 m2/g) and the volume of the
reaction product was found to be 1.2 times larger than that of the original Teflon,
indicating mutually intercalated LiF and carbon. Wide-angle X-ray scattering
data reveal this to be a unique form of polymeric carbon. The degree of
graphitization (six-membered rings) can be increased by heating the films. Low
resolution SEM photographs show the reduced particles to be laminar structures
which decompose into lamellae resembling the known elementary lamellae of
Teflon spherulites. At high resolution, there is some local orientation but the
noduli have pores between them which appear to be formed during LiF
crystallization [36]. A kinetic study of the reduction of Teflon by alkali metal
amalgams has also been reported [37].
If Teflon is held in contact with a platinum wire at a potential more negative
than - 2.0 V in a cell containing a solution of tetrabutylammonium tetrafluoro-
borate in dry dimethyl formamide, it is reduced to a black film, [38]. XPS
analysis of the carbon shows that it is similar to that obtained upon chemical
reduction [39]. A freshly prepared film, in vacuum, has a resistivity of 10-1-10-'
ohm cm which increases rapidly by several orders of magnitude when the film is
exposed to air or protic solvents. Controlled potential electrolysis established
that the reduction requires - 4.1-4.5 electrons per -(C2F4)--- repeat unit, a
value greater than the theoretical value of 4.0 electrons. This is the result of the
formation of a reduced form of carbon similar to the tetra-alkylammonium
intercalation compounds reported earlier [41, 42]. Double-layer capacitance
measurements indicate a surface area of 1000 m2/g, which implies a mean pore
size of 3 nm. Lap shear strengths and contact angle measurements made on
electrochemically reduced Teflon, after various treatments, are shown in Table 3
[42]. The higher lap shear strengths suggest that the surface of the reduced
Teflon has become much more hydrophilic and has become amenable to bonding
with adhesives. The highest lap shear strength occurred for electrochemical
reduction at - 2.5 V in a tetrabutylammonium tetrafluoroborate (TBA TFB)
electrolyte in DMF solvent.
Modification of fluoropolymers by impregnation of carbonyls and subsequent
oxidation [44, 45] produces contact angle lowering, as shown in Table 4,
indicating that the surface free energy has been raised. Doping with tin by
pyrolysis has also been reported [46]. The presence of metal oxides produces a
lowering of the water contact angles. X-ray diffraction measurements of tin-
doped samples gave reflections which suggest a new Sn/Teflon polycrystalline
162

Table 3.
Lap shear strength° and water contact angle data for electrochemically reduced
PTFE [42]

"Composite test pieces consisted of Al/epoxy adhesive/PTFE/epoxy adhe-


sive/AI, tested in tensile shear in accordance with ASTM D 1002-53T. Method
of Sharpe and Schonhorn [43].
'Electrochemically reduced at - 2.5 V in an electrolyte consisting of tetra-
butylammonium tetrafluoroborate (TBA TFB) in DMF.
`Same as b above but in the presence of naphthalene.

Table 4.
Water-PTFE contact angles after Fe(CO)5impregnation followedby oxidation°

° Contact angles measured by the captive bubble method. After two sorption
cycles of Fe(CO)5and two oxidation cycles of 6 h each.

material. These samples are semiconductors with the total conductivity


consisting of contributions from the conduction and impurity bands.
Potassium hydroxide etching produces changes in the hydrophobicity of Teflon
surfaces [47]. The effects are evidenced by lowering of the contact angles of
water and ethylene glycol. SEM measurements indicate no significant surface
roughening, suggesting that the modification is chemical in nature.
Chemical groups arising from the initiator used in the free radical dispersion
polymerization may be present on the Teflon surface [48]. These groups can
markedly affect the adsorption of anionic, cationic, and non-ionic surfactants.
Introduction of -CF3 groups onto the surface of Telfon may be accomplished by
using Surflon®*, a fluorine-containing oligomer composed of a perfluoroacrylate
and co-monomer [49]. Since Surflon selectively exudes on the surface, blending
of small amounts gives large effects. Similarly, blends of poly(methyl-
methacrylate) with poly(vinylidene fluoride) produce a distinct lowering of the
blend critical surface tension with only a few percent of fluoropolymer [50].

*Surflonis the tradename of an oligomeric material marketed by the Asahi Glass Company.
163

3. CRYSTALLIZATIONAND CROSSLINKING BYMETALS

The most direct way to bond a metal to a polymer is by lamination or deposition


of a metal onto the polymer. Intimate contact of the polymer surface by the
deposited metal is necessary for bonding. The normally low surface energy of
fluoropolymers, at room temperature and atmospheric pressure, precludes
intimate contact with metals. An obvious solution to this problem is to laminate
metals to fluoropolymers at elevated temperatures and pressures [51, 52]. It is
also necessary for crystallization of the polymer to occur at the metal-melt
interface and not in the bulk. When crystallization occurs in the bulk, polymer
molecules which cannot be accommodated into the crystal lattice are exuded to
the interface and a weak boundary layer is formed [53].
The joint strength of Teflon, crystallized above its melt temperature, Tm
( Tm=327°C ), was studied by preparing laminates of glass/Au/Teflon/Au/glass,
hot-pressed at 400°C and 17.6 kg/cm2 for 30 min [54]. After removal of the gold ,
with aqueous NaCN, tensile shear test structures were prepared out of Al/
epoxy/Teflon/epoxy/A1 composites, yielding the shear strength results shown in
Table 5. The dramatic effect of crystallization against gold can be seen by the
lowered contact angles of various liquids on gold-crystallized Teflon and
untreated skived Teflon (see Table 6). The weak boundary layer present on
fluoropolymers seems to be a direct function of how the surface is generated.
Evaporation of a metal onto a fluoropolymer surface followed by its removal
provides a modified surface good for bonding other evaporated metals [56]. If
aluminum is evaporated onto Teflon and then removed with NaOH, it provides
an improved surface for evaporation of gold. The tensile shear strengths of
composite structures of Al/epoxy/Au/Teflon/Au/epoxy/Al are shown in Table 7.

Table 5.
Al/Epoxy/PTFE/epoxy/A1 composite tensile shear strengths [54]

°Temperature at which joint was made.


Composites were molded at 400°C, 17.6 kg/cm2 for 30 min.
'CASING (Crosslinking of Activated Species by INert Gas) is a glowdischarge treatment [55].

Table 6.
. Wettability of Au nucleated vs. skived
(untreated) PTFE at 20°C
164

Table 7.
Tensile shear strengths for metal/PTFE/metal joints (56]a°

° Composites were constructed of Al/epoxy/metal/PTFE/metal/


epoxy/Al.
b Resistiveheating evaporation of Al.
'NAOH dissolution of Al followed by resistive heating evapora-
tion of Au.
°Resistive heating evaporation of Au.

X-ray fluorescence of the failed gold/Teflon composites showed that no gold


remained on the Teflon; rather, Teflon was transferred to the gold. This implies
good contact but that the locus of failure was the weak boundary layer present on
the Teflon surface. Lowering of the contact angles of various liquids on the
aluminum-treated Teflon was observed, but SEM studies revealed no apparent
surface roughening. XPS results did suggest, however, an oxygen-containing
hydrocarbon species on the Teflon surface.
Evidence for defluorination of the polymer followed by formation of an
organometallic species [57, 58] has been observed for aluminum evaporated
onto Teflon. It was suggested that the metal fluoride generated by abstraction of
fluorine from Teflon could catalyze a crosslinking reaction of the unsaturated
hydrocarbon species, thus increasing the cohesive strength of the surface region
by removal of the weak boundary layer. Metal fluoride formation at the interface
is shown in the XPS data in Table 8, and fluorine loss is shown in Table 9. It was
noted that the highest lap shear strength obtained on treated Teflon was
produced by the metal that caused the greatest surface composition change.
Surface crosslinking by evaporation of metals is thought to be important for good

Table 8.
XPS data°for Teflon, metallized Teflon, and metal fluorides
[57, 58]

° Estimateduncertainty: t 0.3eV.
bAu 4f7/2'
I
Al 2p.
dTi 2p.
165

Table 9.
F 1 s/C 1 intensity ratios for evapor-
ated metals on Teflon [57, 58]

aNominal 6 nm layer of metal evapor-


ated onto Teflon. I(C ls) is essentially
the same for all the above cases; there-
fore, shielding is not responsible for
change in ratio.

adhesion [59]. Using tensile shear test laminates of Al/epoxy/metal/Teflon/


metal/epoxy/Al, the following order of a series of metals was determined:
Ti > Fe > Ni > A1 > Au > Cu. Although no correlation with shear strength could
be found for contact angle data after metal etching, studies on other polymers
indicate a relationship between the thickness of a crosslinked layer and tensile
shear strength data [59]. The suggestion is that a metal which is more reactive
with Teflon generates a higher crosslink density and thus results in better
adhesion.
Differential scanning calorimetry (DSC) studies of various metals with Teflon
[60, 61] show that thermal events occurring when the mixtures are heated cannot
be assigned to phase transitions. Exotherms appearing in the DSC trace were
correlated with the appearance of metal fluoride peaks in the XPS spectrum.
Similar studies have been carried out with graphite monofluoride [62]. However,
the thermal decomposition of Teflon at elevated temperatures [63] and the
reaction of its decomposition product, tetrafluoroethylene, with metals and metal
oxides was not considered [64].

4. ION AND ELECTRON BEAM MODIFICATION


Beams of ions, molecules, atoms, and electrons have been used to modify fluoro-
polymer surfaces. In many instances, the mechanism of interaction of the particle
with the polymer surface is not well understood. Besides possessing kinetic
energy, a particle approaching a surface may have internal excitation energy as
well as a surface-particle energy of interaction. An ion, for instance, has an
internal characteristic energy typified by its ionization potential. The degree of
surface-particle interaction, and therefore the interaction energy, will obviously
be dependent on the kinetic energy of the particle as well as on the nature of the
particle. An impinging particle may penetrate the surface, be reflected, be
adsorbed, may liberate surface and subsurface species, be charge-neutralized as
it approaches the surface, or may chemically react with the surface. In addition to
the nature of the particle, the intrinsic nature of the species comprising the
surface is important. The final state of a surface after particle bombardment may,
therefore, be dependent on several factors.
Topographical changes in fluoropolymer surfaces have been detected after
166

bombardment with ions, electrons, and neutrals. Pre-sputtering Teflon surfaces


was found to increase the copper-Teflon adhesion [65]. Splitting of the material
into individual filament-like structures following growth of isolated microfissures
was observed after subjecting PTFE to low-keV heavy atom bombardment [66].
A similar topography developed owing to local heating which occurred when a
150 pA, 17 keV electron beam was used. In a study of simulated space environ-
ment, a decrease in the infrared transmittance of an electron-bombarded Teflon
sample was found. Bulk damage to the films was manifested by pinholes, tunnels,
and subsurface channels. Results of SEM analysis indicated evidence for
depolymerization and repolymerization [16]. XPS analyses have show that
defluorination leading to an unsaturated carbonaceous network occurs in PTFE
films exposed to 12.5 keV electrons [67].
An argon ion source used to texture FEP polymer surfaces [68, 69] produced
an improved surface for adherance of sputtered metals. The texturing did not
decrease the surface resistivity, and the sputtered rough metal films had a cone
structure which was more elastic and resistant to mechanical damage than metal
substrate cones. Experiments focusing on bombarding Teflon with different
energy xenon atoms [70] resulted in a model based on microcrack development,
surface splitting, filament formation and subsequent filament-to-cone trans-
formation in the development of surface topography.
The charging phenomenon of electron-beam-irradiated Teflon has been
studied by several researchers [17, 18, 71 ] . A model was developed to study the
effects of various experimental parameters such as beam energy, angle of
incidence, current density, and material thickness and width. It was found that
the charging could be classified as either conduction-limited or emission-limited
for low-beam current densities and high-beam current densities, respectively.
A comparison of electron beam and corona charging was made on (metallized)
Teflon FEP-A foils [72]. It was previously known that charge carriers are
deposited directly onto the surface of FEP after corona charging at room
temperature. By comparison with electron beam charging, it was found that
energetically shallower traps (1.2 eV) are deposited onto the surface and deeper
traps (1.8 eV) are located in the bulk of the material; however, the physical
nature of the traps was not known. The interesting point made in this study was
that no new traps are introduced by electron beam charging. Only bulk traps are
formed by an electron beam, whereas corona charging yields both surface and
bulk traps.
Improvement in bond strengths between gold and Teflon and Teflon FEP-A
has been reported. Irradiation of Teflon with a 1 MeV beam of protons (10"
cm - 2) or helium ions (1013 cm-2) resulted in enhanced adherance of gold films
[73, 74]. In previous studies, a thin mixed layer was predicted to be formed at
the interface; however, Rutherford backscattering (RBS) profiles showed that if a
mixed layer is present it is less than 5 nm thick. Bonding between Teflon FEP
and 50-100 nm evaporated gold was found to be greatly increased if the fluoro-
polymer surface was first subjected to electron beam bombardment. Bond
strengths approaching the bulk strength of Teflon ( - 60 kg/cm2) were obtained
for electron beam energies in the range of 5-20 keV. The gold-Teflon bond
strength for untreated material is - 10 kg/cm2. It was noted that the bond
strength is primarily controlled by surface modifications caused by the electron
167

beam and only to a lesser degree by stored charge. The energy of the injected
particles determines the extent of surface modification, and crosslinking is most
likely to cause an improvement in bond strength.
A grafting technique consisting of irradiation of polymers in methyl acrylate
vapor followed by hydrolysis was used for PTFE and PCTFE [75]. The results of
peel strengths of modified PTFE sheets bonded with an epoxy adhesive are
shown in Table 10. The peel strengths of the grafted PTFE sheets were much
higher than those which were sodium-etched. The bond improvement was
claimed to be due to the mechanical strength of the modified surface layers.
Ion bombardment resulting in texturing of fluoropolymer surfaces yields an
improvement in epoxy bond strengths compared with a sodium-naphthalene
treatment [76]. Large cone or grass-like surface microstructures, parallel to the
direction of the incident ions, result. The degree of crystallinity increases as a
result of this surface structure formation, and the extremely rough surface
morphology enables strong mechanical attachment to high modulus adhesives.
Peel strengths for the various polymers are listed in Table 11. Ion beam texturing
was more effective than sodium-naphthalene treatment for FEP; however, the
opposite was true for PTFE. Similar treatment was used to etch metals, fluoro-
polymers, and chlorofluoropolymers [77]. Textured fluoropolymers produced
greater bond strengths than could be achieved with conventional surface
treatments.
Irradiation of PTFE with MgKa X-rays or 2 keV electrons causes damage
confined to a few tenths of a pm below the fluoropolymer surface [78]. The

Table 10.
Peel strengths of surface-modified PTFE sheets bonded with an epoxy adhesive [75]]

Table 11.
Effects of ion beam texturing vs. sodium-naphthalene treatment on fluoropolymer-epoxy bond
strengths [76].

°Argon ion, 750 eV.


b 0.6mA/cm2.
`Maximum peel strength could not be measured due to cohesive failure of polymer tested.
10.3 mA/cm2.
168

damaged layer was found to be crosslinked or branched PTFE. The usual


decomposition mode of this fluoropolymer is chain scission followed by
unzipping of the polymer, producing a product which is almost 100% monomer.
X-rays were found to produce free F radicals, leaving chains with long-lived
trapped, active sites able to form crosslinking bonds. These active sites, as well as
the F radicals, are available to terminate fragments caused by scission so that
unzipping need not materialize. No monomer was detected in the mass spectrum,
and the effect of electron bombardment was similar, except that the rate of
damage was higher. Both Ni and Au evaporated films adhered better to the
irradiated surface. Ni-F and probable Ni-C chemical interaction, but no Au
interaction, was detected. Similar studies [79] showed epoxy-PTFE joint
strengths of 19.8-25.3 kg/cm2 for electron-bombarded samples compared with 0
kg/cM2 for untreated PTFE. Cohesive failure within the PTFE layer was
= 140.6
characterized by kg/cm2 joint strength. It was also found that electron
bombardment of PTFE in the presence of ND3 resulted in the adsorption of
nitrogen-containing groups at or near the interface. Crosslinking appears to be a
competitive process with nitrogen adsorption, and it was suggested that this
method may be an alternative to plasma processing for modification of polymers
for adhesion improvement. FTIR studies have shown that irradiation of PTFE
with a 6°Co source results in a decrease in sample crystallinity and chain scission.
Production of gaseous products also occurs, as does lowering of the glass
transition temperature [80].
Monte Carlo simulations were performed in order to better understand the
changes in XPS spectra acquired during Ar ion sputtering of PTFE [81]. It was
found that the damaged layer has a homogeneous composition within the
sampling depth of XPS. The spectra could be better explained by assuming a
double fluorine atom elimination mechanism. Bombardment of Teflon and
graphite monofluoride by Xe was found to result in reduction of carbon due to
lattice damage by the impinging ions and implantation of only a small amount of
Xe [82].
If nitrogen or nitrous oxide ions are used to bombard a Teflon surface,
reduction of carbon takes place when fluorine is evolved and the fluorine atoms
are replaced by nitrogen [83]. XPS gives evidence for -CF2- groups
transformed to a -NCF- type. It was suggested that such beam-surface
reactions utilize selectivity and specificity, thus enabling alteration of the
chemical nature of the surface.

5. GLOW DISCHARGE PLASMA MODIFICATION

Glow discharge plasmas may alter the properties of polymers by chemically or


physically inducing changes in the surface or near-surface layers. Such changes
may affect adhesion or wetting properties, molecular weight or crosslinking, and
produce more reactive surface chemical functionalities. Depending on the
reaction conditions, tailoring a polymer surface to specific needs is possible.
Sputter etching of PTFE with argon by RF glow discharge produces micro-
scopic cones and changes the chemistry at the surface [84], resulting in an
improvement of the adhesion properties and wettability. Cones 1 pm in height at
a density of 109 projections/cm2 are produced and -C-O bonds are formed on
169

the surface. The major contribution to changes in wettability was attributed to


physical effects. Both chemical and physical effects contributed to the improved
epoxy bond strength, and treatment times longer than 30 s produced bond
strengths superior to that of sodium-etched PTFE.
Ionizing radiation-induced changes in the chemical composition, crystalline
content and structure, as well as flow properties of PTFE, were studied [84]. It
was found that irradiation in the presence of 02 causes the formation of acid
flouride end groups which may hydrolyze to form carboxylic acid end groups
upon exposure to water vapor. Band assignments used in the infrared analysis
were taken from literature values and are listed in Table 12. The dominant effect
of ionization radiation on PTFE was concluded to be molecular weight degrada-
tion caused by polymer chain scission.

Table 12.
Infrared band assignments for PTFE modified by O, ionizing radiation [85]]

Optimizing the flow rate, plasma power, and treatment time of an acetylene
glow discharge may increase the adhesive bonding of PTFE [85]. A wettable,
rough, crosslinked layer which yields good adhesion can be obtained.
The adhesion of copper to PTFE/glass is improved by an ion plating technique
[86]. If PTFE is sputter-cleaned in a pure noble gas environment, a surface
microroughness favorable for good adhesion of the electron-beam-deposited
copper film is achieved, but the PTFE chains are broken and only weakly
bonded with the inner chains. If a mixture of oxygen with argon or helium is
used, the PTFE chains are broken, but oxygen simultaneously reacts with chain
170

fragments forming volatile compounds. A maximum adhesion strength of 18-


20 x 106 N/M2 for oxygen mixtures was obtained. This is in contrast to a value of
less than < 6 x 106 N/mz for the pure Ar or He discharge.
Polymer weight loss [87] and rate of bond strength increase with plasma
treatment time [88] of fluoropolymers and copolymers were studied. The rate of
detachment of F from the -CFz- chain was found to be less than that of H from
the -CH2- chain, and it was supposed that -CFz-CF2- acts as a blocking
surface. When compared with polyethylene (PE) [89], a longer treatment time
required for PTFE to achieve comparable adhesion was attributed to the
relatively unreactive nature of radicals in PTFE. Rapid defluorination of the
fluoropolymer surface after exposure to H2 in an inductively coupled RF field
was studied by XPS [90], and the formation of an approximately 2 nm boundary
layer of defluorinated polymer was detected.
XPS analyses, coupled with physical determinations of surface properties, can
provide important information on the chemical changes taking place on RF-
sputtered fluoropolymer surfaces [91-93]. In fact, one study shows that the
effects of sputtering are 70% physical and 30% chemical in nature [91]. In a study
of ET-TFE copolymer subjected to He, Ne, Ar, and Kr plasmas, He was found to
be the most efficient gas for crosslinking the outermost few monolayers. Cross-
linking of the subsurface and bulk polymer is most efficient when Ne is used.
Simulated crosslinking studies, compared to XPS of seven carbon-chain model
compounds, suggested that CH2 sites have greater activity than CF2 sites on
exposure to inert gas plasma. Crosslinking and > C=O formation comprise the
early part of the reaction, with the oxidation side reaction dominating as the
exposure time is increased. The XPS ratio of C : F: O for a control FEP film with
textured surface morphology was 1.0:2.18:0 as compared with a smooth,
pinhole-free sputtered film with a ratio of 1.0:0.54:0.38. The analysis confirmed
that the outermost layers of the sputtered specimen are partially oxidized and
that the oxygen atoms are capable of reorienting themselves either from the
surface or to the bulk, depending on the nature of their environment.

6. SUMMARY
A variety of methods used to modify fluoropolymer surfaces has been described
in this review. Table 13 summarizes these methods and lists the resulting changes
in fluoropolymer surface properties. It is easily seen that a lack of information
relating surface property changes to adhesion exists. In many instances where
such data are reported, the methods used to evaluate adhesion are varied and, as
such, correlation of adhesion to surface property changes upon modification is
wanting. Until such a correlation is made, the ultimate choice for a fluoro-
polymer surface modification scheme will depend on such factors as the type of
interface to be considered and the complexity or availability of treatment.

REFERENCES
1. A. J. Kinloch, J. Mater. Sci. 15,2141 (1980).
2. J. E. E. Baglin, in: Ion Beam Modification of Insulators, P. Mazzoldi and G. W. Arnold (Eds),
Chap. 15, Elsevier, Amsterdam (1986).
3. M. Brenman and Ch. H. Lerchenthal,Polym. Eng. Sci. 16, 747(1976).
171
172

4. H. V. Boenig (Ed.), Proc. Annu. Int. Conf Plasma Chem. Tech., vol. 2, p. 17. Technomic,
Lancaster, PA (1986).
5. H. Schonhorn, H. L. Frisch and G. L. Gaines, Jr., Polym. Eng. Sci. 17, 440 (1977).
6. D. M. Brewis and D. Briggs, Polymer 22, 7-15 (1981).
7. E. B. Hale, W. J. James, A. K. Sharma and H. K. Yasuda, in: Proc. Int. Conf Modification of
Surface Properties, Metal Ion Implantation, V. Ashworth, W. A. Grant, R. P. M. Procter (Eds),
vol. 3, p. 167. Oxford Press, New York (1982).
8, G. Surendran, W. J. James, W. Brearley and E. B. Hale, J. Appl. Polym. Sci.: Appl. Polym. Symp.
38, 75 (1984).
9. H. Yasuda, H. C. Marsh, E. S. Brandt and C. N. Reilley, J. Polym. Sci.: Polym. Chem. Ed. 15,
991 (1977).
10. C. Weaver,J. Vac.Sci. Technol. 12, 18-25 (1975).
11. L. J. Gerenser, J. Vac.Sci. Technol.A6, 2897 (1988).
12. D. H. Buckley, Mater. Res. Soc. Symp. Proc. 40, 359(1985).
13. D. T. Clark,Pure Appl. Chem. 54,415-438 (1982).
14. K. L. Mittal, J. Adhesion Sci. Technol. 1, 247-259 (1987).
15. H. F. Mark and N. G. Gaylord (Eds), Encyclopedia of Polymer Science and Technology,vol. 13.
John Wiley,New York (1970).
16. O. Berolo, Proc. Int. Symp. Spacecraft Materials in Space Environment, Toulouse, France (June
1982).
17. K. G. Balmain and W. Hirt, IEEE Trans. Nucl. Sci. 27, 1770-1775 (1980).
18. R. D. Reeves and K. G. Balmain, IEEE Trans. Nucl. Sci. 28, 4547-4552 (1981).
19. E. Sacher and J. R. Susko, J. Appl. Polym. Sci. 27, 3893 (1982).
20. R. H. Dahm, in: Surface Analysis and Pretreatment of Plastics and Metals, D. M. Brewis (Ed.),
p. 227. Macmillan, New York (1982).
21. E. R. Nelson, T. J. Kilduff and A. A. Benderly, Ind. Eng. Chem. 50, 329-330 (1958).
22. R. J. Purvis and W. R. Beck, U.S. Patent 2,789,063 (16 April 1957 to Minnesota Mining and
Manufacturing Co.).
23. G. Rappaport, U.S. Patent 2,809,130 (8 October 1957 to General Motors Corp).
24. D. W. Dwight and W.M. Riggs,J. Colloid Interface Sci. 47, 650-660 (1974).
25. R. R. Rye and J. A. Kelber, Appl. Surf. Sci. 29, 397-410 (1987).
26. C. A. Costello and T. J. McCarthy, Macromolecules 17, 2940-2942 (1984).
27. Z. Iqbal, D. M. Ivory, J. S. Szobota, R. L. Elsenbaumer and R. H. Baughman. Macromolecules
19, 2992-2996 (1986).
28. V. Faingold-Murshak and B. Karvaly, Polym. Commun. 26, 358-359 (1985).
29. A. J. Dias and T. J. McCarthy, Macromolecules 18, 1826-1829 (1985).
30. A. J. Dias and T. J. McCarthy,Macromolecules20, 2068-2076 (1987).
31. C. A. Costello and T. J. McCarthy, Macromolecules20, 2819-2828 (1987).
32. D. D. MacNicoland C. D. Robertson, Nature 332, 59-61 (1988).
33. J. Jansta, F. P. Dousek and V. Patzelova, Carbon 13, 377-380 (1975).
34. J. Jansta and F. P. Dousek, Electrochim. Acta 18, 673-674 (1973).
35. F. P. Dousek, J. Jansta and J. Baldrian, Carbon 18, 13-20 (1980).
36. Z. Pelzbauer, J. Baldrian, J. Jansta and F. P. Dousek, Carbon 17, 317-322 (1979).
37. F. P. Dousek and J. Jansta, Electrochim. Acta 20,1-6 (1975).
38. D. J. Barker, D. M. Brewis, R. H. Dahm and L. R. J. Hoy, Electrochim. Acta 23, 1107-1110
(1978).
39. H. Brecht, F. Mayer and H. Binder, Angew.Makromol. Chem. 33, 89 (1973).
40. J. Simonet and H. Lund, J. Electroanal. Chem. 75, 719-730 (1977).
41. J. O. Besenhard and H. P. Fritz, J. Electroanal. Chem. 53, 329-333 (1974).
42. D. M. Brewis, R. H. Dahm and M. B. Konieczko, Angew. Makromol. Chem. 43, 191-194
(1975).
43. Sharpe and H. Schonhorn, Adv. Chem. Ser. 43, 189 (1964).
44. R. Baumhardt-Neto, S. E. Galembeck, I. Joekes and F. Galembeck, J. Polym. Sci., Polym. Chem.
Ed. 19, 819-829 (1981).
45. F. Galembeck, J. Polym. Sci., Polym. Chem. Ed. 16, 3015-3017 (1978).
46. F. Froissart and A. J. Horodecki, Thin Solid Films 75, 119-124 (1981).
47. Y. Rotenberg, S. Srinivasan, E. I. Vargha-Butler and A. W. Neumann, J. Electroanal. Chem. 213,
43-51 (1986).
173

48. H. E. Bee, R. H. Ottewill, D. G. Rance and R. A. Richardson, in: Adsorption from Solution, R. H.
Ottewill, C. H. Rochester and A. L. Smith (Eds), pp. 155-171. Academic Press, London (1983).
49. S. Ohtoshi, Jpn Plastics Age 27-30 (Jan.-Feb. 1985).
50. T. Davidson, Polym. Prepr. Am. Chem. Soc. Div. Polym. Chem. 19,175-179 (1978).
51. H. Schonhorn, H. L. Frisch and T. K. Kwei,J. Appl. Phys.37, 4967-4973 (1966).
52. H. Schonhorn, Macromolecules 1, 145-151 (1968).
53. H. D. Keith and F. J. Padden, Jr., J. Appl. Phys. 35,1270-1285 (1964).
54. H. Schonhorn and F. W. Ryan, J. Adhesion 1, 43-47 (1969).
55. H. Schonhorn, F. W.Ryan and R. H. Hansen, J. Adhesion 2, 93-99 (1970).
56. R. F. Roberts, F. W. Ryan, H. Schonhorn, G. M. Sessler and J. E. West, J. Appl. Polym. Sci. 20,
255-265 (1976).
57. H. Schonhorn and R. F. Roberts, Polym. Prepr. Am. Chem. Soc. Div. Polym. Chem. 16, 146
(1975).
58. M. S. Toy, J. Polym. Sci., Part C. 34, 273-279 (1971).
59. S. L. Vogeland H. Schonhorn, J. Appl. Polym. Sci . 23, 495-501 (1979).
60. P. Chadman and G. M. Gossedge, J. Mater. Sci. 14, 2672-2678 (1979).
61. G. Pocock and P. Cadman, Wear 3 7,129-141 (1976).
62 P. Cadman and G. M. Gossedge, J. Mater. Sci. 14, 1465-1472 (1979).
63. J. C. Siegle,L. T. Muus, T.-P.Lin and H. A. Larsen, J. Polym. Sci., Part A 2, 391 (1964).
64. W.A. Brainard and D. H. Buckley, Wear 26,75-93 (1973).
65. C.-A. Chang, Appl. Phys. Lett. 51, 1236-1239 (1987).
66. R. Michael and D. Stulik, Nucl. Instrum. Methods Phys. Res. B14, 278-281 (1986).
67. D. T. Clark and W. J.Brennan, J. Electron Spectrosc. Relat. Phenoin. 41, 399-410 (1986).
68. M. J. Mirtich and J. S. Sovey, US Patent 4,199,650 (22 April 1980).
69. M. J. Mirtich and J. S. Sovey, NASA, TM-73778 (June 1978).
70. R. Michael and D. Stulik, J. Vac.Sci. Technol.A4, 1861-1865 (1986).
71. R. C. Hazleton, R. J. Churchill and E. J. Yadlowsky, IEEE Trans. Nucl. Sci. 26, 5141-5145
(1981).
72. H. von Seggern,J. Appl. Phys. 52, 4086 (1981).
73. B. T. Werner, T. Vreeland, Jr., M. H. Mendenhall, Y. Qui and T. A. Tombrello, Thin Solid Films
104, 163-166 (1983).
74. G. M. Sessler, J. E. West, F. W. Ryan and H. Schonhorn, J. Appl. Polym. Sci. 17, 3199-3209
(1973).
75. S. Yamakawa,Macromolecules12, 1222-1227 (1979).
76 B. A. Banks, J. S. Sovey,T. B. Miller and K. S. Crandall, NASA, TM-7888 (June 1978).
77. M. J. Mirtich and J. S. Sovey, NASA, TM-79004 (Dec. 1978).
78. D. R. Wheeler and S. V. Pepper, NASA, TM-83413 (Aug. 1983).
79. J. A. Kelber, J. W.Rogers, Jr. and S. J. Ward, J. Mater. Res. 1, 717 (1986).
80. H. Vanni and J. F. Rabolt, J. Polym. Sci.: Polym. Phys. Ed. 18, 587-596 (1980).
81. T. Takahagi and A. Ishitani, Macromolecules20, 404 (1987).
82. J. A. Taylor, G. M. Lancaster and J. W. Rablais, Appl. Surf. Sci. 7, 503-514 (1978).
83. J. A. Taylor, G. M. Lancaster and J. W. Rablais, J. Am. Chem. Soc. 100, 4441-4447 (1978).
84. S. Yamamoto, H. Tabata, M. Ezoe, K. Uemori, Y. Oya and T. Moriuchi, Oyo Botsuri 53, 639
(1984).
85. W. K. Fisher and J. C. Corelli, J. Polym. Sci.: Polym. Chem. Ed. 19, 2465-2493 (1981).
86. A. Moshonov and Y.Avny, J. Appl. Polym. Sci. 25, 771-778 (1980).
87. A. Celerier and J. Machet, Thin Solid Films 148, 323-332 (1987).
88. T. Hirotsu and S. Ohnishi, J. Adhesion 11, 57-67 (1980).
89 C. A. L. Westerdahl, J. R. Hall and D. W. Levi, Polym. Prepr. Am. Chem. Soc. Diu: Polym.
Chem. 19, 538-543 (1978).
90. H. Schonhorn and R. H. Hansen, J. Appl. Polym. Sci. 11, 1461-1474 (1967).
91. D. T. Clark and D. R. Hutton, J. Polym. Sci.: Polym. Chem. Ed. 25, 2643-2664 (1987).
92. T. Moriuchi, S. Yamamoto, M. Ezoe, H. Tabata, K. Uemori, E. Shigeta, Y. Ohya and A. Tsmura,
Proc. 7th Int. Vac.Congr.2, 1501-1504 (1977).
93. D. T. Clark and A. Dilks, J. Polym. Sci.: Polym. Chem. Ed. 16, 911-936 (1978).
94. E. Ruckenstein and S. V. Gourisankar, J. Colloid Interface Sci. 107, 488-502 (1985).

You might also like