Sunteți pe pagina 1din 9

DK4036_book.

fm Page 1 Monday, April 25, 2005 12:18 PM

5
The Theory of Adhesion

5.1 Contact Angle Equilibrium ................................................5-1


5.2 Forces of Attraction ............................................................5-3
Carl A. Dahlquist 5.3 Real and Ideal Adhesive Bond Strengths ...........................5-8
3M Company References .......................................................................................5-9

When pressure-sensitive adhesive is applied to a smooth surface, it sticks immediately. The application
pressure can be very slight, not more than the pressure due to the weight of the tape itself. The adhesive
is said to wet the surface, and, indeed, if the tape is applied to clear glass and one views the attached
area through the glass, it is found that in certain areas the adhesiveglass interface looks like a liquidglass
interface. From this one would infer that a pressure-sensitive adhesive, even though it is a soft, highly
compliant solid, also has liquidlike characteristics. Some knowledge of the interaction between liquids
and solids is benecial to the understanding of adhesion.

5.1 Contact Angle Equilibrium


When a drop of liquid is placed on a surface of a solid that is smooth, planar, and level, the liquid either
spreads out to a thin surface lm, or it forms a sessile droplet on the surface. The droplet has a nite
angle of contact (Figure 5.1). The magnitude of the contact angle depends on the force of attraction
between the solid and the liquid and the surface tension of the liquid. The contact angle equilibrium has
received a great deal of attention, principally because it is perhaps the simplest direct experimental
approach to the thermodynamic work of adhesion.
Many years ago Young1 proposed that the contact angle represents the vectorial balance of three tensors,
the surface tension of the solid in air (sa), the surface tension of the liquid in equilibrium with the vapor
(lv), and the interfacial tension between the solid and the liquid (sl), The force balance can be written

sa = lv cos ++ sl (5.1)

Youngs equation has come under criticism on the grounds that the surface tension of a solid is ill
dened, but most surface chemists nd his equation acceptable on theoretical grounds.
The equation can be written as a force equilibrium or as an energy equilibrium, because the surface
tension, expressed as a force per unit of length, will require an energy expenditure of the same numerical
value when it acts to generate a unit area of new surface.
Harkins and Livingston2 recognized that Youngs equation must be corrected when the exposed surface
of the solid carries an adsorbed lm of the liquids vapor. The solidair-plus-vapor tensor, sv, is less
than the solidair tensor, sa. Harkins and Livingston introduced a term, e, to indicate the reduction thus:

sv = sa e (5.2)

5-1

2006 by Taylor & Francis Group, LLC


DK4036_book.fm Page 2 Monday, April 25, 2005 12:18 PM

5-2 Coatings Technology Handbook, Third Edition

Air

Liquid

v
l
sa sl

Solid

FIGURE 5.1 Contact angle equilibrium of surface and interfacial tensions.

and the modied Young equation becomes accordingly

sa = lv cos + sl++ e (5.3)

The work of adhesion (Wa) is the thermodynamic work necessary to separate the liquid from the
solid without performing any additional work, such as viscous or elastic deformation of either the liquid
or the solid. Dupr3 is credited with the denition of the work of adhesion. When a liquid is separated
from a solid, new solidair and liquidair interfaces are created, and solidliquid interface is destroyed
(Figure 5.2).
The Dupr equation becomes

Wa = lv++ sa sl (5.4)

Combining Equations 5.4 and 5.3 yields.

Wa = lv (1 ++ cos) ++ e (5.5)

Under some circumstances, e is negligible, and Equation 5.5 simply reduces to the original
YoungDupr equation:

Wa = lv (1 ++ cos) (5.6)

The work of adhesion can then be obtained by measuring the surface tension of the liquid and the
contact angle; sa need not be known.
Zisman and his associates4,5 at the Naval Research Laboratories have done extensive and careful
investigation of the contact angles of liquids on solids. They found that when the surface tensions of a
family of liquids were plotted against the cosines of their contact angles on a given solid, the data fell,
with some scatter, along a relatively straight line. The surface tension value at which this line intersected
the ordinate [i.e., where cos = 1.00 (or = 0)] was designated the critical surface tension, c.
Some have equated c of a solid to its surface tension or surface energy, but this was never done or
condoned by Zisman. The critical surface tension of a solid is simply the highest surface tension among
the surface tensions of liquids that will spread on the solid. It is, of course, a measure of the attractive
force a liquid experiences when it comes in contact with the solid.

Liquid
lv
sl (or sv)
sa
Solid

FIGURE 5.2 Wa = lv + sa sl.

2006 by Taylor & Francis Group, LLC


DK4036_book.fm Page 3 Monday, April 25, 2005 12:18 PM

The Theory of Adhesion 5-3

70

WA ergs/cm2 60

50

40
0 20 40 60 80 100
Contact Angle

FIGURE 5.3 Work of adhesion of liquids on a surface having a c of 21.5 ergs/cm2.

Some have postulated that a contact angle of 0 is essential for the formation of a strong adhesive
bond. This is said to be requisite for wetting. Wetting, however, is a relative term. Surely the molecules
of a liquid and a solid will be in close contact at the interface even though the contact angle is greater
than 0. In fact, if one measures the works of adhesion between a family of liquids and a solid, one nds,
as illustrated in Figure 5.3, that the work of adhesion can maximize at a relatively high contact angle.
Thus, spreading is not requisite for good adhesion on thermodynamic grounds, but it may be highly
desirable in real situations for maximizing contact area and minimizing interfacial aws and defects.

5.2 Forces of Attraction


All atoms or molecules, when brought into sufciently close proximity, exhibit attraction for each other
as a result of asymmetric charge distributions in their electron clouds. Under special circumstances, if
there are xed electrical charge distributions (e.g., in electrets) or xed magnetic polarization, repulsion
may occur, but these are special and rare cases. Repulsion also occurs when matter is compressed to the
point that the outer electron orbits begin to interpenetrate each other; but when electrical charges and
magnetic domains are randomly distributed, and the liquids and solids are not under extremely high
compressive forces, the forces of attraction prevail.
The forces take several forms. The forces due to the uctuation in the electron cloud density are
universal. These forces were rst recognized by Fritz London6 and are sometimes called the London
forces, but more frequently, they are called dispersion forces. The uctuation in electron density
produces temporary electrical dipoles, which give rise to universal attraction in matter. These are the
forces that account for cohesion and adhesion in aliphatic hydrocarbons. They are relatively weak atomic
interactions, but because all the atoms are involved, the cumulative effect is great. The dispersion forces
depend markedly on the distance of separation, decreasing as the seventh power of the separation between
isolated atom pairs, and as the third or fourth power of the separation between planar surfaces.
The dispersion forces are common to all matter and can account for universal adhesion in condensed
media, but other forces can contribute substantially to adhesion in specic cases. There may be attractive
forces between permanent dipoles, where the potential energy of head-to-tail interaction of two dipoles
having dipole moments 1 and 2 is given in the lowest energy conguration, as

2006 by Taylor & Francis Group, LLC


DK4036_book.fm Page 4 Monday, April 25, 2005 12:18 PM

5-4 Coatings Technology Handbook, Third Edition

21 2
U ntP =
r3

where r is the center-to-center distance between the dipoles.


If the rotational energy is less than the thermal energy of the system, then

212 12
U K = Keesom potential =
3kTr 6

where k is Boltzmanns constant (0.0821 1atm/mol deg), and T is absolute temperature (K).
There may be dipole-induced dipoles, where the potential energy of interaction is given by

12 2 + 22 1
U1 =
r6

where 2 and 1 are the molecular polarizabilities.


There may be acidbase interactions7,8 across the interface that can lead to strong bonding. Examples
are hydrogen bonding, Lewis acidbase interactions, and Brnsted-type acidbase interactions.
Covalent bonding between adhesive and adherend, if achievable either by chemical reactions or by
high energy radiation, can lead to very strong bonds.
Interdiffusion, usually not achievable except between selected polymers, can also lead to high adhesion.
The force of attraction between planar surfaces has been derived from quantum mechanical consid-
erations by Casimir, Polder,9 and Lifshitz.10
Lifshitz calculated the attractive forces between nonmetallic solids at distances of separation sufciently
large that the phase lag due to the nite velocity of electromagnetic waves becomes a factor. He obtained
the following relationship between the attractive force and the known physical constants:

2 hc (e o 1)[(e o )]
F=
240d 4 (e o + 1)

where F is the attractive force per unit of area, h is Plancks constant, C is the velocity of light, d is the
distance of separation, eo is the dielectric constant, and (eo) is a multiplying factor that depends on the
dielectric constant as follows:

1/eo: 0 0.025 0.10 0.25 0.50 1


(eo): 1 0.53 0.41 0.37 0.35 0.35

Strictly speaking, the dielectric constant in this expression should be measured at electron orbital
frequency, about 1015 Hz. However, if we assume handbook values of the dielectric constant at 106 Hz,
which for nylon, polyethylene, and polytetrauoroethylene, are 3.5, 2.3, and 2.0, respectively, the corre-
sponding (eo) values are 0.37, 0.36, and 0.35. The force values then stand in the ratios 0.11 to 0.056 to
0.039. When normalized to F (nylon) = 1.0, they fall to the following ratios:

F(nylon), = 1.00; F(PE), 0.51; F(PTFE), 0.35

When the c values (dynes/cm) of these three materials are similarly normalized to the c values for
nylon, the values fall in remarkedly similar ratios.

Nylon PE PTFE
c 56 31 18.5
c(norm) 1.00 0.55 0.33

2006 by Taylor & Francis Group, LLC


DK4036_book.fm Page 5 Monday, April 25, 2005 12:18 PM

The Theory of Adhesion 5-5

250

1
200 F
D4.07
Separation in Angstrom Units

100 1
F
D2.94

50

20
1 10 100 1000
Force Constant

FIGURE 5.4 Attraction between ideally planar solids.

In the Lifshitz equation, the force of attraction is shown to decrease as the inverse fourth power of the
distance of separation. However, when the separation becomes so small that the phase lag in the inter-
action no longer is signicant (it is of the order of 6 at a separation of 50 ), the attractive force varies
as the inverse third power of the distance of separation. This has been veried experimentally,11 although
the direct measurement is extremely difcult (Figure 5.4). The forces existing at separations greater than
50 contribute very little to adhesion.
Some 30 years ago Good and Girifalco reexamined the interfacial tensions between dissimilar liquids
and developed a theory of adhesion.12 They found that the work of adhesion, given by

Wa = L1 + L2 L1L2

could be approximated quite well by the geometric mean of the works of cohesion of the two liquids
when the only attractive forces of cohesion are dispersion forces:

Wa = 2( L1 L2 )1/2

However, in some liquid pairs (e.g., water and hydrocarbons), this did not hold, and they coined an
interaction parameter, , given by

L1 + L2 L1L2
=
2( L1 L2 )1/2

Thus,

Wa = 2( L1 L2 )1/2

2006 by Taylor & Francis Group, LLC


DK4036_book.fm Page 6 Monday, April 25, 2005 12:18 PM

5-6 Coatings Technology Handbook, Third Edition

For water on a parafnic hydrocarbon, where the contact angle is 108, would have a value of about
0.55. For hexadecane on polyethylene, is very near unity. Good and his associates11,12 have provided
directions for calculating , and they give experimental and calculated values for several combinations
of water and organic liquids.
Fowkes13 approached the problem from a different point of view. He reasoned that the only forces
operable at the interface between water and an aliphatic hydrocarbon molecule contain no hydrogen
bonding groups and no xed dipoles.
Fowkes also assumed that the work of adhesion would be given by twice the geometric mean of the
surface energies of the two liquids on either side of the interface but now taking into consideration only
the dispersion force components of the surface energies. For the work of adhesion between water (L1)
and n-octane (L2), we have

Wa = 2( DL1 DL2 )1/2 = L1 L1L2

where the superscript D stands for the dispersion energy component of the total surface energy. Accepted
values for the surface energies and interfacial energies are as follows:

L1 = 72.8 ergs/cm 2 ; L2 = DL2 = 21.8 ergs/cm 2 ; L1L2 = 50.8 ergs/cm 2

If these values are substituted into the equation above to solve for DH 2O , we get 22.0 ergs/cm2. Fowkes
evaluated several water-aliphatic hydrocarbon systems and found that they all yielded essentially the same
value for the dispersion energy component of the surface energy of water, 21.8 0.7 ergs/cm2.
Turning now to the work of adhesion and the interfacial energy between mercury and aliphatic
hydrocarbon, Fowkes calculated the dispersion energy component of the surface energy of mercury. Using
n-octane as the hydrocarbon liquid having a surface energy of 21.8 ergs/cm2 (all of it attributed to
dispersion forces), the surface energy of mercury, 484 ergs/cm2, and the interfacial energy, 375 ergs/cm2,
we have

Wa = 2( DHg nDoct )1/2 = Hg + noct ( Hg ,noct)

Wa = 2( DHg 21.8)1/2 = 484 + 21.8 375

DHg = 196.2

The average DHg for a series of mercuryaliphatic hydrocarbon systems yielded 200 7 ergs/cm2 for
the dispersion energy component of the surface energy of mercury.
Since the remaining forces that contribute to the surface energy of mercury are metallic forces, the
only interacting forces at the watermercury interface are the dispersion forces, and the work of adhesion
is given by

Wa = 2(200 21.8)1/2 = 484 + 72.8 ( H g , H 2O )

from which

( H g , H 2O ) = 424.7 ergs/cm 2

This compares very favorably with the measured value of 426 ergs/cm2.

2006 by Taylor & Francis Group, LLC


DK4036_book.fm Page 7 Monday, April 25, 2005 12:18 PM

The Theory of Adhesion 5-7

The work of adhesion due to dispersion forces is numerically small in work or energy units. For
example, the work of adhesion of methylene iodide on polyethylene is 82 ergs/cm2 ( = 52). This
small value is not, however, indicative of a small force of attraction across the interface. Keep in mind
that the work is the product of force and displacement, and that the attractive force, at separation
distances less than 50 (5 107 cm) increases as the inverse of displacement raised to the third
power (Figure 5.4).
The molecules at the interface are at an equilibrium distance of separation where attractive forces and
repulsive forces balance. The variation in the repulsive forces with distance of separation has a dependence
several orders of magnitude higher than the attractive forces (of the order of 1012 for atom pairs and 108
for repulsion forces across a hypothetical plane). We can calculate the maximum force of attraction by
equating the work of adhesion to the work of separation.
Let Fa indicate the attractive force, Fr the repulsive force, x the distance separation, and d the equilibrium
distance. We cannot measure d directly, but we can estimate it from calculations of the distance between
molecular centers in a liquid of known specic gravity and molecular weight. In the case of methylene
iodide (sp g 3.325, mol 267.9), we calculate the separation to be about 5 108 cm between the centers
of adjacent molecules.
If we take 5 108 cm as a reasonable distance of separation across the interface between methylene
iodide and polyethylene, and we accept the force versus distance relationships for attraction (a) and
repulsion (r), we can write:

3
d
Fa = ( Fa )e
x
8
d
Fr = ( Fr )e
x

where the subscript e stands for equilibrium. At equilibrium we have the condition that (Fa)e = (Fr)e.
We can then express the work of adhesion as

3 8
d d
Wa =

d
( Fe ) dx
x
d
( Fe ) dx
x

The solution is

d d
Wa = Fe
2 7

For methylene iodide on polyethylene, Wa is 82 ergs/cm2. Taking d as 5 108 cm, Fe = F = Fr = 4.92


109 dynes/cm2.
The maximum attractive force is encountered where the difference between the attractive forces and
the repulsive forces maximizes as separation proceeds. This occurs where (d/x)3 (d/x)8 maximizes, at
about x = 1.22d.
At this displacement, F = 0.347Fe, or, in the case of methylene iodide and polyethylene, at 1.71 109
dynes/cm2 (about 25,000 psi). This would be the maximum attractive force experienced when separation
of the materials is attempted; it far exceeds the average stresses that are typically observed when adhesive
bonds are broken.
Others have calculated theoretical forces of adhesion by other approaches. All yield results that predict
breaking strength far exceeding the measured breaking strengths.

2006 by Taylor & Francis Group, LLC


DK4036_book.fm Page 8 Monday, April 25, 2005 12:18 PM

5-8 Coatings Technology Handbook, Third Edition

25

20

Failure Stress, Dyne/cm2 107


15

10

0
0 0.1 0.2 0.3 0.4
Thickness of Adhesive Layer, cm

FIGURE 5.5 Adhesive layer thickness and strength of butt joints.

5.3 Real and Ideal Adhesive Bond Strengths


How, then, does one account for the fact that theoretical or ideal bond strengths do not seem to be
attainable? The measured cohesive strengths of solids also fall far short of theoretical values. Only whisker
crystals of silicon, graphite, iron, and the like have measured tensile strengths approaching their theoretical
tensile strengths.
J. J. Bikerman contended that all adhesive bonds are awed and contain weak boundary layers, and
hence, always fall short of their theoretical strengths. Undoubtedly this is often the case; but aws, at
least gross aws, can be minimized, and fracture in weak boundary layers, which can be detected, is not
always observed.
All destructive tests, whether done in tension, shear, or peel, involve stress concentrations. Unless the
stress is uniformly distributed over a very small area, as in the tensile strength tests on the whiskers,
there will be localized stresses that far exceed the average stress. It is well known that the measured average
breaking strengths of adhesive bonds broken in tension decrease as the adhesive thickness increases
(Figure 5.5), and it can be shown that tensile stresses at the periphery of the bond are higher than interior
stresses. If a aw leads to highly localized stress, cleavage may be initiated and proceed catastrophically.
Bonds pulled in shear not only experience shearing stress concentrations but also tearing stresses.14,15
Peeling is a deliberate application of stress concentration.
Deformation and ow also contribute to stress concentrations and failure. It would be advantageous
to match the mechanical properties of the adhesive to the mechanical properties of the adherend, but
this is rarely feasible. Instead, the adhesive designer resorts to the expedient of toughening the adhesive
and incorporating materials that arrest crack propagation, thereby maximizing the work necessary to
destroy the bond.
In many instances, the adhesiveadherend interface is more accurately described as an interphase.
A case in point is the bonding of aluminum to aluminum with structural adhesive. The surface of
aluminum is really aluminum oxide, which, depending on the manner in which it was formed, will vary
in strength and porosity. The adhesive penetrates and locks into the oxide lm, and the bond strength
can be greatly enhanced by use of a surface treatment that produces a strong, well-bonded oxide layer.

2006 by Taylor & Francis Group, LLC


DK4036_book.fm Page 9 Monday, April 25, 2005 12:18 PM

The Theory of Adhesion 5-9

Adhesiveadherend bond strengths are often enhanced by priming. A classic example is the bonding
of vulcanized rubber to steel, in which the steel is rst electroplated with a thin coat of copper, and the
rubber compound is cured on the copper surface under heat and pressure. It is believed that the sulfur
in the vulcanizate bridges to the copper by chemical bonding to give a strong bond, and the copper, in
turn, is strongly bonded to the steel.
Priming is often used on bonding adhesives to plastic lms. For example, the rst transparent pressure-
sensitive tape, which comprised a cellophane lm and a natural rubber rosin adhesive, would undergo
separation of the adhesive from the lm under humid conditions. The problem was solved by rst
applying to the cellophane a thin prime coat, a blend of natural rubber and casein, then coating the
adhesive over the primer.
Surfaces notoriously difcult to bond to, such as polyethylene, polypropylene, and Teon, are modied
by treatments that make the surfaces more polar and possible chemically active, for example, by corona,
plasma, or chemical treatments. These treatments may also remove weak boundary layers.
Though strong, durable adhesive bonds are usually the goal of adhesive technology, there is also a
need for bonds that are deliberately made weak. This need arises in the pressure-sensitive tape industry,
where it is desirable to have tape that unwinds easily from the roll, and especially where pressure-sensitive
adhesives are to be transferred from a carrier lm to another surface. The surfaces that provide easy
release typically have low critical surface energies, almost totally dominated by the dispersion energy
component. Also, for release coatings to function well, there must be no mutual solubility between them
and the adhesives. Silicone release coatings, which consist mainly of polydimethyl siloxane, provide the
easiest release. They have low critical surface energies, though not as low as certain uorocarbon polymers.
They also have a high degree of incompatibility with the pressure-sensitive adhesives that release well
from them, but these criteria alone do not explain the low level of adhesion. In addition, they differ from
other release coatings by being soft and elastic rather than hard. This feature may serve to enhance the
stress concentration when the adhesive is separated from the release liner.
In conclusion, adhesive bond strengths measured by destructive tests will never approach theoretical
value, but the intrinsic attractive forces can be manipulated by the choice of materials and surface
treatments to produce a wide range of practical bond strengths.

References
1. T. Young, Phil. Trans. R. Soc. London, 95, 65 (1805).
2. W. D. Harkins and H. K. Livingston, J. Chem. Phys., 10, 342 (1942).
3. A. Dupr, Thorie Mcanique de la Chaleur, Paris, 1869, p. 393.
4. W. A. Zisman, Ind. Eng. Chem., 55, 18 (1963).
5. E. G. Shafrin, in Polymer Handbook, J. Brandrup and E. M. Immergut, Eds. New York: Wiley-
Interscience, 1966, pp. 111113.
6. F. London, Trans. Faraday Soc., 33, 8 (1936).
7. R. S. Drago, L. B. Parr, and C. S. Chamberlain, J. Am. Chem. Soc., 99, 3203 (1977).
8. F. M. Fowkes, J. Adhes. Sci. Technol., 1, 7 (1987).
9. H. B. C. Casimir and D. Polder, Phys. Rev., 73, 360 (1948).
10. E. M. Lifshitz, C. R. Acad. Sci. USSR, 97, 643 (1954).
11. D. Tabor and R. N. S. Winterton, Nature, 219, 1120 (1968).
12. L. A. Girifalco and R. J. Good, J. Phys. Chem., 61, 904 (1957).
13. F. M. Fowkes, J. Phys. Chem., 66, 382 (1962).
14. O. Volkersen, Luftfahrforschung, 15, 41 (1938).
15. M. Goland and E. Reissner, J. Appl. Mech., Trans. ASME, 66, 17 (1944).

2006 by Taylor & Francis Group, LLC

S-ar putea să vă placă și