Stability of epoxy/glass interfaces

advertisement

Lehigh University

Lehigh Preserve

Theses and Dissertations

2003

Stability of epoxy/glass interfaces

Orasa Khayankarn

Lehigh University

Follow this and additional works at: http://preserve.lehigh.edu/etd

Recommended Citation

Khayankarn, Orasa, "Stability of epoxy/glass interfaces" (2003).

Theses and Dissertations.

Paper 789.

This Thesis is brought to you for free and open access by Lehigh Preserve. It has been accepted for inclusion in Theses and Dissertations by an authorized administrator of Lehigh Preserve. For more information, please contact preserve@lehigh.edu

.

Khayankarn, Orasa

Stability of

Epoxy/Glass

Interfaces

May 2003

Stability of Epoxy/Glass Interfaces

by

Orasa Khayankarn

A Thesis

Present to the Graduate and Research Committee of Lehigh University

In Candidacy for the Degree of

Master of Science in

Polymer Science and Engineering

Lehigh University

May, 2002

Acknowledgements

After 2 years at Lehigh University, many people have had a great influence and assistance in the accomplishing of my goals. First of all, I would like to express my great appreciation and respect to my advisor, Dr. Raymond A. Pearson, who has been patient, understanding, encouraging and very helpful to me since my first day working under his supervision. Without his guidance, I would have never completed this work. Second, I would like to express my appreciation to the Royal Thai Government for all of their

/\ financial support. Without them, I would have never been here.

This work could not be achieved without the help from Karl I.

Loh and Zymet

Incorporated for their financial support to this project and the Dow Chemical Company for their thermoplastic adhesion promoters. I also would like to thank Arlan Benscoter for teaching me how to grind and polish specimens, Al Miller for providing XPS data and

Donna Narsavage-Heald for teaching me many of the techniques used in this work and her helpful idea and suggestions.

Finally, I would like to express a very special appreciation to my parents for their great support and inspiration. Without them, none of this would have been possible.

111

TABLE OF CONTENTS

Content

Cover Page

Certificate of Approval

Acknowledgements

Page i ii

111

Table of Contents iv

List of Tables ....

~.

... ........... .. . . ... ... ..... .. .. ..... .. .. .. ............. ... ... ... .... .. ... vii

.

0 fF'

19ures ,

...

Vlll

Abstract '"

CHAPTER 1 INTRODUCTION

CHAPTER 2 BACKGROUND

2.1 Fracture Mechanics

2.2 Hygrothermal Aging.

1

2

7

13

2.2.1 Diffusion Behavior '" 15

2.2.1.1 Theory of Diffusion in Polymeric Materials .......... 16

2.2.1.2 Moisture Absorption in Epoxy Resins

2.2.2 Irreversible Effect of Hygrothermal Aging

18

22

2.3 Adhesion Promoters

2.3.1 Effect of the Coupling Agent Layer Thickness

2.3.2 Mechanism of Adhesion Promotion '"

2.3.2.1 Chemical Bond Theory

2.3.2.2 Deformable Layer Theory

25

27

29

29

30

IV

2.3.2.3 Surface Wettability Theory

2.3.2.4 Restrained Layer Theory

2.3.2.5 Reversible Hydrolytic Bond Theory

2.3.2.6 Oxide Reinforcement

2.3.2.7 Other Mechanisms

2.3.3 Types of Adhesion Promoters

2.3.3.1 Silanes

2.3.3.2 Titanates

2.3.3.3 Organic Resin Adhesion Promoters

2.3.4 Previous Work of Epoxy/Glass Stability

2.3.5 Effects Other than Adhesion Promotion

2.4 Objectives i

CHAPTER 3 EXPERIMENTAL APPROACH

3.1 Materials

3.1.1 Epoxy Model: Bisphenol F Resin with

2,4.EMI Curing Agent

3.1.2 Adhesion Promoters

3.1.2.1 Monomeric Adhesion Promoters

3.1.2.2 Polymeric Adhesion Promoters

3.1.3 Borosilicate Glass

3.2 Sample Preparations......................................................

3.2.1 Surface Treatment of Adhesion Promoters

48

48

55

56

56

52

52

53

32

34

35

40

30

30

31

32

44

45

46

46 v

3.2.2 Double Cantilever Beam (DCB) Specimen.................... 57

3.3 Double Cantilever Beam (DCB) Testing

3.4 Evaluation of two Silane Adhesion Promoters

3.4.1 Effect of Aging Time on G le

3.4.2 Moisture Uptake of Aging Specimen

3.5 Evaluation of Polymeric Adhesion Promoters

60

61

61

61

62

CHAPTER 4 RESULTS AND DISCUSSION

4.1 Evaluation of two Silane Adhesion Promoters '" 63

4.1.1 Effect of Aging Time on Strain Energy Release Rate (G le)

66

4.1.2 Moisture Uptake of Aging Specimen 68

: 4.2 Evaluation of Polymeric Adhesion Promoters

4.2.1 PHAE Series Polymeric Adhesion Promoters

73

75

4.2.2 BLOX®200 Series Polymeric Adhesion Promoters 78

CHAPTER 5 CONCLUSIONS

References

Appendix 1 X-ray photoelectron spectroscopy (XPS) analysis of borosilicate glass

Appendix 2 Percent failure and G le of DCB specimens treated by polymeric adhesion promoters

Appendix 3 Study of subcritical crack growth of epoxy/glass interfaces

Vita

'"

81

" 83

91

95

97

101 vi

LIST OF TABLES

TABLE

Table 2.1: Plot of primary bond formation for various types of coupling agents

Table 2.2: Typical silane adhesion promoters commercially available

Table 2.3: Recommended silane adhesion promoters for various resins '"

Table2.4: Effect of substrate on silane adhesion effectiveness , ;

Table2.5: Typical titanate coupling agent

Table 3.1: PHAE epoxy based thermoplastic adhesion promoters in different solvent systems

Table 3.2: BLOX® epoxy based thermoplastic adhesion promoters in different solvent systems

Table 3.3: Composition (%wt) of borosilacate glass

Table A1.1: XPS analysis of borosilicate glass surface

Table A1.2: Composition of borosilicate glass surface

Table A1.3: XPS analysis of bulk borosilicate glass

Table A1.4: XPS analysis of bulk borosilicate glass

Table A2.1: Percent failure and

Ole of DCB specimens treated by

PHAE series polymeric adhesion promoters

Table A2.1: Percent failure and

Ole of DCB specimens treated by

BLOX®200 series polymeric adhesion promoters

PAGE

34

55

96

38

39

40

43

54

56

92

92

95

91

91

Vll

LIST OF FIGURES

FIGURE NUMBER

Figure 1.1: Schematic diagram of typical forming operation for glass fibers

Figure 2.1: Mechanical adhesion is a function of both surface energy and

PAGE

3

7 substrate energy dissipation.

Figure 2.2: the double cantilever beam (DCB) adhesive joint.

9

Figure 2.3: Typical load, P, versus displacement, 8, trace for DCB specimen. " 10

Figure 2.4: Fracture stress of epoxy/mild steel butt joints as a function of the time exposed to various environments.

14

Figure 2.5: The percentage of water content versus the ratio of square root of ...... 20

,time to the thickness of sample for aging at 65°C in (a) water and

(b) a 5% (w/w) water solution of NaCl.

Figure 2.6: Model of diffusion front near the substrate/polymer transition

Figure 2.7: Hydrolysis reaction of epoxy resin

Figure 2.8: Effect of hygrothermal aging on the tan 8 versUs temperature relationship.

Figure 2.9: Adhesion promoters provide a strong interphase region

Figure 2.10: Schematic diagram of the polymer-coupling agent-substrate

Figure 2.11: Reaction of silane adhesion promoters on substrate

Figure 2.12: Reaction of silane with polymer :

Figure 2.13: Titanate condensation on a hydroxyl containing surface

26

26

, 36

37

41

22

23

24 viii

Figure 3.1: Diglycidyl ether of bisphenol F.

Figure 3.2: 2-ethyl-4-methyl-imidazole

48

.48

Figure 3.3: A reaction mechanism sequence of 2,4-EM! cured Bisphenol F. ., 50

Figure 3.4: y-Aminopropyltriethoxysilane

Figure 3.5: 2-(3,4-Epoxycyclohexyl)ethyltrimethoxysilane

Figure 3.6: The top and bottom glass plates ofDCB specimen

Figure 3.7: Double cantilever beam (DCB) specimen

Figure 3.8: A diagram of DCB specimen preparation

Figure 3.9: Typical double cantilever beam (DCB) analysis.

Figure 4.1: Percent failure ofDCB specimens under 85°C/85% RH. aging

52

52

58

58

59

60

64 condition using adhesion promoter as a surface treatment.

Figure 4.2: Percent failure of DCB specimens under 85°C/85% RH. aging condition using adhesion promoter as an additive in epoxy resin

Figure 4.3: Gte drop of untreated, y-APS and ECH treated DCB specimens

65

66 under 85°C/85% RH. aging condition

Figure 4.4: Water content (%) as a function of aging time of untreated, y-APS ..... 68 and ECH treated DCB specimens under 85°C/85% RH.

aging condition

Figure 4.5: Water content (%) as a function of the square root of aging time of .... 69 untreated, y-APS and ECH treated DCB specimens under

85°C/85% RH. aging condition ix

Figure 4.6: Relation between

Gtc drop and water content ofDCB specimens under 85°C/85% RH. aging condition of (a) untreated samples,

71

(b) y-APS treated samples and (c) ECH treated samples

Figure fl..7: Percent failure and G tc of DCB specimens of untreated and y-APS treated samples

74

Figure 4.8: Percent failure and G tc of DCB specimens treated by

PHAE polymeric adhesion promoter in different solvent systems

Figure 4.9: Gtc ofDCB specimens treated by PHAE series polymeric adhesion promoter in different solvent systems at aging time of

(a) 7 days and (b) 14 days

Figure 4.10: Percent failure and G tc ofDCB specimens treated by

BLOX®200 series polymeric adhe.sion promoter in different solvent systems

Figure 4.11: G tc of DCB specimens treated by BLOX®200 series polymeric adhesion promoter in different solvent systems at

75

77

78

80 aging time of (a) 7 days and (b) 14 days

Figure A1.1: XPS spectrum of borosilicate glass surface 93

94 Figure A1.2: XPS spectrum of bulk borosilicate glass

Figure A3.1: Double Cantilever Beam (DCB) specim~n , 97

Figure A3.2: (a) curve relationship between load and time; (b) curve relationship.. 98 between crack growth rate and strain energy release rate (G) at crack tip x

Figure A3.1: Plot of log crack growth rate (daldt) in units of (m/s) vs strain energy release rate (G) in units of J/m

2 of untreated sample

99

100 Figure A3.2: Plot of log crack growth rate (daldt) ys. strain energy release rate (G) of untreated sample xi

ABSTRACT

Stability of epoxy/glass interfaces under hygroythermal aging was studied by applying a fracture mechanics approach.

The use of two silane adhesion promoters, yaminopropyltriethoxysilane (y-APS) and 2-(3,4-epoxycyclohexyl)ethyltrimethoxysilane

(ECH), on strain energy release rate, G le, was also studied. As expected, the G le dropped with aging time which can be explained by loss of interfacial forces. Adhesive strength improvement can be seen in the y-APS and ECH treated specimens both before and after aging due to the introduction of chemical bonding at the interfaces. Chemical bonding, which is able to resist to the hygrothermal condition, was formed in the interphase region. Moreover, the results show that applying adhesion promoters on glass surface retard water absorption of the specimens leading to adhesive strength improvement. Two-stage water absorption was observed. The first step was a Fickian type related to the process of saturation of epoxy matrix and at longer exposure time and the The second was possibly osmotic pressure driven influenced by amount of water at the interfaces-

Furthermore, the wet adhesive strength in terms of G le for commercial polymeric adhesion promoters, the PHAE series and the BLOX®200 series, under hygrothermal aging was investigated.

It was found that 2% PHAE in Dowanol® PM, hydroxy group containing solvent, was the best system in the PHAE series. For the BOLX®200 series, 2% BLOX/2%

MAH showed higher adhesive strength than 2% BLOX/l % MAR. The higher MAR content provides more active sites reacting with either epoxy resin or glass surface. In conclusion,

PHAE and BLOX®200 series improve stability of epoxy/glass interfaces under hygrothermal conditions; however, y-APS still out performs both types of the~oplastic adhesion promoters.

1

CHAPTERl

INTRODUCTION

Epoxy resins are widely used as matrices for adhesives, organic coatings, and structural composite materials due to their superior mechanical properties [1-4] However, it is well known that these properties are strongly affected by moisture absorption, which causes plasticization [5~6], lowering the glass transition temperature [5-7], and diminishing their mechanical strength [7,8]. Identification of the basic phenomena responsible for glass/epoxy adhesion has proven necessary in both the composite and adhesives industries.

Fowkes et al. [9] has reviewed that the adhesion of polymer to inorganic oxides such as glass is very dependent on acid-base interactions between the acidic or basic surface sites of the glass and the basic or acidic functional sites of the polymer. This interaction is quite strong in dry condition but the strength of adhesion drop dramatically when expose to severe condition involving high temperature and high humidity.

It is generally thought that adhesion is improved by forming chemical bonds between the adherend and adhesive.

Adhesion promoters with dual functionality are widely used to join two materials by the formation of chemical bonds between polymer and substrate. Chemical bonding can produce durable interfaces, even under severe humidity and thermal conditions.

Adhesion promoters used to improve the hygrothermal stability of reinforc~d composites improve the adhesion of the matrix resin to the reinforcement material and aid in the transfer of stress during the process of mechanical deformation. The typical glass fiber

2

formation is shown in Figure 1.1. This high quality.glass fiber is usually dawn from hightemperature platinum alloy bushing tip at high cooling rates and linear speeds into a water mist. This process generates a uriique composition and structure in top 0.5-5 nm of the glass fiber surface [10]. Molecular dynamics simulations on silica predict that a high concentration of silanol groups (Si-OH) will be generated on surfaces created in the presence of water [11]. Within milliseconds of solidifying, the fiber pass across an applicator roll where they are coated with sizing solutions that promote strength, stiffness, and durability in final composite.

~~~..,.....:r;;..y..,.......,........~

~I-Bushing Tip

Temperature> 1000°C

Water

Spray ---

==.....a~~.I..I-.,o~~:--

Temperature < 100°C

Sizing - - - .

Applied '""""-.........

i'TT'iTT"i'ri-.........

..J

ollII

Glass Fibers

Figure 1.1: Schematic diagram of typical forming operation for glass fibers.

Many researchers have investigated the mechanical properties of glass fiber or glass bead reinforced epoxy composites. Failure or debonding has been observed at the epoxy/glass interfaces when they are exposed to the hot/humid environments. The extent

3

of interfacial failure increased with time of exposure. It has been repeatably shown that adhesion promoters on glass surface improves mechanical properties under hygrothennal conditions. Improvement in interfacial strength between glass and epoxy has been attributed to the use of a coupling agent [12,13], although a greater advantage of the adhesion promoter was the increase in the adhesive bond's resistance to moisture attack.

It has been seen that epoxy will readily bond to clean glass, but the interface is highly susceptible to hydrolysis. Apart from improving the durability of adhesion between polymers and inorganic reinforcements, organosilanes have been shown to prevent cure inhibition by mineral fillers. In epoxy, it was observed that silane treatments of fillers overcome cure inhibition as measured by cure exotherms [14,15].

According to Dwight et al. [16], several studies have suggested that silanes form primary chemical bonds with both glass substrate and polymeric resins [17]. Although there is a large amount of indirect evidence supporting the theory of surface. bond formation, direct evidence is somewhat limited [18]. Miller and Ishida have cited a few referenced that, at the time, reflected the majority of the direct spectroscopic evidence for presence of adsorbate-surface covalent bonds. For instance, Miller and Ishida observed the formation of a covalent bonds between organofunctional trialkoxysilane adsorbate molecules and a lead-oxide surface using diffuse reflectance infrared spectroscopy

[18,19]. The resulting surface bond was the plumbosiloxane linkage (pb-O-Si). Culler

[18,20] monitored the adsorption of y-aminopropyltriethoxysilane (y-APS) from aqueous solution on silicon oxide using Fourier transform infrared (FUR) spectroscopy. It was

4

observed that the bands in the region 1200-1000 cm-

1 from Si-Q.:-Si bonds had shifted in frequency as a function of concentration in solution.

Spectroscopic studies by Chiang and Koening [15,21] focused on the interaction between the silane coupling agent and the polymeric resin system. They observed that aminofunctional silanes were unsuitable for reinforcements in anhydride-cured epoxy composites, since the anhydride had reacted with primary amine groups of silane to form cyclic imides. Although complete reaction of epoxy and anhydride was obtained, some of the potential crosslinking in the interphase region was lost [15,21].

It is interesting to understand the effect of environmental attack on adhesive strength at glass/epoxy interfaces. As a result, this study has focused the effort on elucidating glass/epoxy adhesion loss under hygrothermal aging. The adhesive strength has been investigated by using a fracture mechanics approach. The results are given in term of strain energy release rate (G\c) and the effect of moisture on the drop of G\c will be highlighted.

The common commercial adhesion promoters can be divided into two types: monomeric adhesion promoters and polymeric adhesion promoters. Many publications have reported adhesion improvement when using various monomeric adhesion promoters, such as silanes and titanates. However, Bell et al. [22] have studied the durability of the polymer-metal interface region when a polymeric adhesion promoter was employed to enhance bonding. They pointed out some advantages of polymeric adhesion promoters. First, the usual schematic diagram of an adhesion promoter shows a relatively low molecular weight molecule with one end attached to the metal oxide and

5

the other to the polymer above it. However, in practice it is difficult to obtain a unifonn monomolecular layer, then the weak link in the strength chain is often between the layers in the adhesion promoters. This is a fundamental disadvantage of monomeric adhesion promoters. Silane adhesion promoters avoid this to a considerable degree by polymerizing on the surface and forming a network within the multilayer, resulting in a structure capable of load transmission.

Another advantage of polymeric adhesion promoters is that they have the ability to absorb a portion of the mechanical and thennal stresses generated in the interphase region which arise due to the mismatch between the moduli and thennal expansion coefficients of the substrate and the polymer. Bare interfaces will have difficulty withstanding the stresses developed upon external mechanical forces, thennal cycling, or water absorption. A third advantage of polymeric adhesion promoters is that the ratio of hydrophobic to hydrophilic components can be controlled; the composition is not fixed, but can be adjusted to suit particular needs. As, mention above, it is very interesting to investigate the changes in adhesive strength of both monomeric and polymeric adhesion promoters in epoxy/glass interface system under hygrothermal aging conditions. The use of polymeric adhesion promoters is a rather new field and the use of polymercapton esters has resulted in same success.

6

CHAPTER 2

BACKGROUND

2.1 Fracture Mechanics

Mechanical adhesion is the result of interactions occurring at the substrate adhesive interface as well as defonnation mechanisms that occur away from the interface in either the adhesive and/or the adherent (Figure 2.1) [23]. Therefore, to successfully understand adhesion issues, a combination of surface science and fracture mechanics is essential in order for significant progress to be made.

yl----.,fa

Y 2 - -

1iI .......I - - -

Y12 y

Process zone

Adhesive

Process zone

Figure 2.1: Mechanical adhesion is a function of both surface energy and substrate energy dissipation.

7

Surfaces can be characterized in terms of surface tensions using contact angle measurements as shown by Lloyd [24]. In turn, surface tensions can be related to the thermodynamic work of adhesion by the following equation.

(2.1)

Where W12 is the thennodynamic work of adhesion (Helmholtz free energy per unit area) in unit of mJ/m

2

, Y1 denotes the surface tension (surface free energy) of substrate 1,

Y2 denotes the surface tension of substrate 2, and

Y12 denotes the interfacial surface energy.

In the absence of excessive plastic deformation in either substrate, it is convenient to relate surface free energies to a fracture mechanics parameter such as a critical strain energy release rate, G c• which is in units of J/m

2

For a homogeneous material this relationship can be expressed as follows:

(2.2) where YP is plastic energy dissipation term.

For real adhesives that debond at the adhesive/adherent interface there are often viscoelastic/plastic processes occurring in the adhesives which must be accounted for.

Gent [25] has proposed the following relationship between the interfacial strain energy release rate,

Gcinter and the thermodynamic work of adhesion W

12

(2.3) where ~ is a term related to viscoelastic/plastic losses at the crack tip. Therefore, to

8

) predict adhesion one must focus on both the thermodynamic work of adhesion as well as loss processes that occur in the adhesive and the adherent near the crack tip.

The concepts of continuum fracture mechanics have been widely employed in studies concerned with crack growth in adhesive joints [26]. Of the various test specimens that have been used to measure the critical strain energy release rate, G e, the double-cantilever beam (DCB) specimen shown schematically in Figure 2.2, has been one of the most popular.

P

/

Bonded end-blocks

Initial crack Adhesive layer

P

Figure 2.2: the double cantilever beam (DCB) adhesive joint

Assuming that the specimen behaves in a linear-elastic manner upon loading, there are four different linear-elastic fracture-mechanics (LEFM) methods for analyzing the data contained in the load-displacement traces.

First, the "area" method, where for elastic behavior the value of G c may be defined by

G = I1U c Bl1a

(2.4) where B is the width of the DCB specimen and, for example, I1U\ is the area under the

9

load-displacement trace and !:J.a

is the increase in crack length from al to a2, as shown in

Figure 2.3. Now when the loading and unloading relations are linear, the LEFM approach may be used and, for example

(2.5) where PI and ~h are the load and displacement, respectively, at a crack length al, and P2 and 8

2 are the restively values at a crack length a2.

4 0 0 . , . - - - - - - - - - - - - - - - - - - - - - ,

300

100

0 - - 1 " " " ' - - - - , . - - - - , . - - - - . , . - - - - - - - - , - - - - - 1 o

20 40 60 80 100

Displacement (mm)

Figure 2.3: Typical load, P, versus displacement, 8, trace for DeB specimen

Secondly, in the "compliance" method, from Irwin and Kies [27], the value of the adhesive fracture energy, G e, from an LEFM test is given by

10

p

2 dC

G = - c

2B da where P is the load and C is the compliance, given by

C=oIP

(2.6)

(2.7) where 0 is the displacement corresponding to a load P.

To evaluate G c via the

"compliance" method, a plot of C versus the crack length, a, may be constructed. The plot of C against a is then curve-fitted using an appropriate polynomial function, and differentiated. Then, knowing the values of the load, P, and the differential, dC/da, at a given crack length, the value of G c as a function of the crack length may be evaluated using Equation 2.6.

From simple beam theory, a third approach to analyzing the data is possible, since the value of the compliance, C, is given by o

2a

3

C = - = -

P 3EJ

Now, whenE s

» En, then I ';::!Bh

3

/12 and

(2.8)

(2.9) where E s the flexural modulus of the fiber-composite substrate arms, which are now assumed to act as simple beams, and E a is the modulus of adhesive layer, I is the second moment of area and h is the thickness of a fiber-composite arm of the DeB specimen,

Hence, from Equation 2.6 and 2.8

11

p 2 a 2

G = -

C

BE I s which may also be expressed as

(2.10)

(2.11)

-.

Fourthly, the "displacement" method, where substitution for P from Equation 2.8

into Equation 2.10 yields

G =3PO

C

2Ba

(2.12)

Now as long as the load-displacement relationships for the specimen are linear for loading and unloading at any given crack length, which is true for all of the present work, then any ~f the four methods described above should give identical results. Furthermore, another cross-check on the analysis methods is provided by calculating the value of the modulus, E s, of the substrate arms. This is given by

(2.13) and may obviously be compared with the value obtained by direct experimental measurements of a beam of the substrate.

The methods suggested here are formulated using methods similar to those postulated by Griffith [28]. Griffith hypothesized that all real "elastic" bodies have inherent cracks in them. He hypothesized that a quantity of energy to make the most critical of these cracks grow would need to come from the strain energy in the body and work applied by loads. Conservation of energy dictates that a crack can grow only when

12

the strain energy released as the crack grows is sufficient to account for the energy required to create the new "fracture" surface. In Griffith's original work, he considered only perfectly elastic systems. He conducted his confinnation experiments on glass, which behaves as a nearly ideal elastic material. In this case the fracture energy is very closely associated with the chemical surface energy.

Indeed, Griffith was able to establish such a correlation by measuring the surface tension of glass melts and extrapolating back to room temperature. Most engineering solids are not purely elastic and the energy required to make a crack grow involves much more than just the chemical surface energy.

In fact, the energy dissipated by other means often dominates the process and may be several orders of magnitude higher than the chemical surface energy. As explained in modem texts on fracture mechanics, this has not prevented the use of fracture mechanics for analysis of fracture in other quasi-elastic systems, where typically the other dissipation mechanisms are lumped into the fracture energy term G c.

Techniques are also being developed to use the basic approach for systems that experience extensive viscoelestic and plastic deformation.

.2.2 Hygrothennal Aging

The effects of environmental aging, especially at high temperatures and high humidity (hygrothennal aging), on the physical properties of the resin are associated with moisture-induced chemical modification, plasticization, and/or micromechanical damage

[29]. One of the most important requirements of an adhesive joint is the ability to retain a significant proportion of its load-bearing capability for long periods under the wide variety of environmental conditions which are likely to be encountered during its service

13

life. In the case of attack in the interfacial regions, then, water is the most damaging environment. In practice in Northern Europe climates the critical concentration of water translates to requiring a relative humidity above about 60 to 80% for attack to be observed. However, if s~ch bonded structures are to be exported to a hot/wet climate, e.g.

the Far East, then it is easy for water to attack the bonds. These points are illustrated in

Figure 2.4. [30] where, as may be seen, the adhesive joints exposed to 55% R.H. at 20°C do not lose any strength, unlike those immersed in water. Increasing the temperature of the environment increases the rate of strength loss.

-

CIS

~ t>

...

30

40

.a

""'

""' cu t.l

CIS

J=

112

Ul cu

20

:g,

10

~ tl

=

0

20°C; 55 % R.H.

0

20°C; H

2

O

••

A

40°C; H

2

O

60°C; H

2

O

---

90°C; H

2

0 - -

.

1:1

500 1000 1500

Time in environment (h)

2000 c:r

2500

Figure 2.4: Fracture stress of epoxy/mild steel butt joints as a function of the time exposed to various environments.

14

In many applications, which often involve some of the most critical uses of

, adhesive technology, the interfaces are subjected to an environment which also happens to be one of the most potentially damaging. Indeed, it is probably true to state that the most important challenges are (i) to develop adhesive systems which possess excellent long-term durability, especially systems acceptable to the general engineering industries, and (ii) to develop test methods and models to predict the service life from short-term experiments, and thereby convince the potential user that an adequate durability will be realized.

2.2.1 Diffusion Behavior

Polymer based-materials are exposed to humid environments, where water molecules, as well as other low molecular weight substances, can easily migrate in the polymeric matrix, even at ambient temperatures, and modify their physical properties.

Moisture transport in polymer system is related to the availability of molecular sized holes in polymer structure and polymer-water affinity [31-34]. The availabity of the holes depends on the polymer microstructure, morphology, and crosslink density, which are functional of degree of cure, stoichiometry, molecular chain stiffness, and the cohesive energy density of the polymer [32]. Polymer-water affinity is related to the presence of hydrogen bonding sites along the polymer chains, which create attractive forces between the polymer and water molecules. The water molecules that are free to move through the holes, or free volume, are often referred to as unbound molecules. Because this unbound moisture is filling free volume, it does not cause dimensional changes of polymer. In contrast, water molecules that attach to the polymer chain via hydrogen bonding, referred

15

to as bonded molecules, disrupt the interchain hydrogen bonding including swelling, and plasticizing the polymer [32-34].

2.2.1.1 Theory of Diffusion in Polymeric Materials

The classical limiting case of diffusion can be described using Fick's second law with a constant diffusion coefficient of diffusivity, D [35] such that

(2.14) where c is the moisture concentration and t is time. For one-dimensional diffusion through an infinite plate of thickness L, this equation reduces to [36]

(2.15) where D x is the diffusivity through the thickness of the material. The diffusion of moisture into glassy polymers generally follows this law when this moisture transport is completely controlled by diffusion, that is, when diffusion mechanisms related to degradation, molecular relaxation, or insufficient curing, for example, are not active. [37]

Further, D x often exhibits a dependence on temperature that follows an Arrehenius relationship [37] given by

(2.16) where Do is a constant coefficient, Ao is the activation energy for diffusion, R is the universal gas constant, and T is the temperature measured on an absolute scale.

16

For most pnictical applications, changes in the percentage of moisture content, M, can be measur~d gravimetrically, that is, as changes in the weight or mass of the polYmer specimen [38] as in

(2.17) where

Wt is the weight or mass of the material at any time, t, and

Wd is the weight or mass of the dry material. The total weight or mass of the moisture in the specimen at any time is given by

(2.18)

For an initial dry material, the value of D x can be calculated from the initial slope, F, of a plot of M(t) as a function of t

1/2 fL using the following relation [39]:

D x

=(!!-J(~J sat

(2.19) where M sat is the equilibrium moisture saturation level for a particular environment. Such a plot should be linear up to M = 0.6 M sat with less than 2% deviation for true Fickian diffusion. [39]

Although the diffusivity does not, in general, depend on the moisture concentration of the surrounding environment, the moisture saturation level is often observed to depend on relative humidity according to the relationship [38]

(2.20)

17

where a and b are constant parameters, and RH is the percentage of relative humidity.

M sat has been observed to be both dependent [40,41] and independent [39] of temperature for different polymer systems. These seemingly contrasting observations have been explained through thermodynamic arguments. [40] For water in epoxy, the heat of solution, Mis, is negative. This factor is opposed thermodynamically by the heat of vaporization of water, Ml y.

If

Mis

=

-My then M sat will not be a function of temperature.

For some systems, this equality does not hold and a temperature-dependent saturation level is observed.

2.2.1.2 Moisture Absorption in Epoxy Resins

In the case of epoxy, significant amount of free volume exist, even at temperatures 50-150 a

C below glass transition temperature, T g, that are often used in moisture absorption studies.

Complex expressions, which take into account parameters, such as the formation of hydrogen bonds, [42,43,44,45] the heterogeneity of the epoxy resin system, [46,47] changes in free volume, [48] relaxation phenomena during hydration, [49,50] degradation of epoxide material, [51] the depression ofT g,

[52] and the formation of clusters, crazes, and micro-cracks during hydration, have been proposed by many authors. [53,54]

The physical aspects of the transport process are also of great importance.

Adamson [55] has postulated that the transport of moisture below T g is a three-stage process in which the absorbed moisture first occupies the free volume.

In the second

18

stage, water becomes bound to network sites, causing swelling. Finally, water enters the densely crosslinked regions.

Yang et al. [56] suggest that while the driving force for the absorption of water is the electrostatic attraction between water and the functional groups of the resin

(especially hydroxyls produced during curing reactions or amine groups when there is an excess' of the curing agent), the equilibrium degree of water uptake is determined primarily by unoccupied volume ofthe epoxy resin.

Carfagna et al. [57] also indicate that water sorption in polYmers containing potential hydrogen bonding sites is driven by strong interactions between water molecules and particular segments or groups present in the polYmer. .

Ivanova et al. [58] studied the combined effects of elevated temperature and moisture on the dicyandiamide cured digylcidyl ether of bisphenol A (DGEBAIDICY) epoxy resin. The material was aged in deionized water and 5% (w/w) NaCl solution at

65°C. Two stage absorption was observed with deionized water, as seen below in Figure

2.5 (a), and the quansiequilibrium-water content was independent of the hygrothermal history. The equilibrium weight after dehydration depended on the hygrothermal history due to the presence of irreversibly absorbed water and leaching of material. The first stage of absorption was found to be Fickian and was interpreted as being related to the proc~ss of saturation of epoxy matrix. The driving force for the second stage was the osmotic pressure, and it was not observed when the material was aged in NaCI solution, as shown in Figure 2.5 (b). The water caused irreversible damage to the resin through microcavity formation, and part of the water was moleculary dispersed in the epoxy

19

matrix and part resided in microcavities.

It was observed that the longer the exposure, the greater the amount of residual water. IR results showed that the hygrothermal aging in salt solution led to chemical modification of the resin and removal of excessive hardener from the sample.

8,.---------------,

500 1000 1500 lIZ lIZ ·1

TIme lThickne&& (s mm)

2000

(a)

2500

!

..

~ lP

~

0

(.)2

== 1

4

5

500 1000 1500 lIZ lIZ .1

Time 1TI'Ilckn8&& (a mm)

2000

(b)

2500

Figure 2.5: The percentage of water content versus the ratio of square root of time to the thickness of sample for aging at 65°C in (a) water and (b) a 5% (w/w) water solution ofNaCl. The symbol (t.), (.) and (0) represent the samples exposed to water with hydration time of 74, 163 and 7250 hours, respectively and the symbol (0), (.) and (0) represent the samples exposed to 5% (w/w) NaCI

Solution with hydration time of74, 152 and 7700 hours, respectively.

20

Moreover, hydration-dehydration cycling led to an increase in the diffusion coefficient of the first stage of the rehydration process and increased the rate of the second stage because of progressive cavity formation.

Ivanova and co-worker [58] also investigated the effect of hydration-dehydration cycling on the dynamic mechanical responses ofDGEBA-DICY epoxy resin. They found that no effect of the hygrothermal history on the dynamic mechanical properties was apparent and they are a function only of the water content. Water absorbed at weight fraction up to 4.3% caused plasticization of the polymer matrix, and the measured decrease in the glass transition temperature could be interpreted in terms of the polymerdiluent theory, Water absorbed at level >4.3% weight fraction resided in microcavities formed as result of the aging process and did not further affect the dynamic mechanical response of the resin. The dynamic mechanical properties in the presence of NaCl suggested that microcavity growth was prevented and the amount of water absorbed did not exceed 4.1

% weight fraction.

The comparison of diffusion of water into an epoxy adhesive between bulk and adhesive joint has been reported by Zanni-Deffarges and Shanahan [59]. In the bulk, diffusion behavior was found to be reasonably approximated by the Fick model and as diffusion progress, the elastic modulus decreased. In case of adhesive joint, it has been shown that the elastic modulus deceased much more quickly. Therefore, they inferred that water is entering by another route, as well as by conventional diffusion within the polymer.

It seems to be that water enters the system by seepage close to the interface or in the interphase region at a faster rate and may then, in turn, diffused towards the bulk of

21

the glue line, as seen in Figure 2.6. Clearly the seepage process may play an important role by also provoking the failure of interfacial bonding.

Substrate

(1) y12

"Wet" adhesive

(2) y13

"Dry" adhesive

~

(3)

Water diffusion

Figure 2.6: Model of diffusion front near the substrate/polymer transition. Terms y12, y23 and y13 represent interfacial 'tension' between the substrate, 'wet' and

'dry' adhesive.

2.2.2 Irreversible Effect ofHygrothermal Aging

Irreversible interactions between water and a model epoxy resin, diglycidyl ether of bisphenol A/dicyandiamide (DGEBAlDICY), during hygrothermal aging was investigated by Xiao et al. [60]. The results clearly showed that irreversible interactions occurred between water and the resin at elevated temperature. Water can lead to the incorporation of C-O and N-CO-N groups in the resin and cut the epoxy blackbone chain.

It is believed that the tertiary amine groups in the cured resin may well be the weakest points, reacting with water via hydrolysis and causing dissociated chain segments by scission, as shown in Figure 2.7. The reaction of waters with the tertiary amine groups in the resin forms N-R and -OR bonds in the polymer chains. The reaction products might

22

undergo further reactions either with each other or with the crosslink chains or can leach out of the bulk resin.

---.. R-NH-C N + HO-R

Figure 2.7: Hydrolysis reaction of epoxy resin.

.Interestingly, subsequent studies by Xiao and Shanahan [61] also showed irreversible effects of hygrotherrnal aging on DGEBSIDICY epoxy resin. The results showed that water immersion at 50, 70 and 90°C led to the introduction of carbonyl groups in the resin and chain scission of the crosslink structure. These irreversible structure changes induced the addition of water to polymer initially and weight loss of the material eventually. They also led to the irreversible decrease of the glass transition temperature of DGEBAIDICY after hygrothermal. As seen for Figure 2.8 below, the tan

8 versus temperature curve for the aged material does not recover its initial form after complete drying of the polymer.

It is known that, in the epoxy resin, the glass transition temperature could be related to the crosslink density. With a decrease of crosslink density, the glass transition shifts to a lower temperature. Therefore, the DMTA results also suggests the cleavage of crosslinks during hygrothermal aging.

23

o

50 100 150 200 250

Figure 2.8: Effect of hygrothermal aging on the tan 8 versus temperature relationship:

(a) before aging; (b) after aging in 90 a

C water for 15 days; (c) after aging in

90 a

C water for 15 days, then complete drying.

Xiao and Shanahan [62] also investigated the swelling behavior ofDGEBAIDICY epoxy resin during hygrothermal aging. The results showed that the rate of swelling of polymer is less than that attributable to the mass of water absorbed initially; however, the rates. equivalised later. They had also found that the swelling process was not fully reversible. When the aged specimens were dried asymptotically, their volume were larger than their corresponding volumes before aging by a value of ~V/Vo of 0.01 where ~V and

Vo are the volume change and the initial dry volume, respectively. The results suggested that hygrothermal aging irreversibly changed the molecular structure of the resin.

24

2.3 Adhesion Promoters

Adhesive failure is a critical problem in the application of adhesives, surface coatings and polymer-filler composites. Many surface preparation techniques have been employed to achieve high initial adhesion; these range from removal of surface contamination (solvent and vapor degreasing) to changes in substrate profile (grit blasting) to chemical modification (phosphating of steel, anodic treatments of aluminum).

None of these methods solve the most critical problem in adhesion technology: that of the damaging effect of water on organic/inorganic bonds.

It has been shown by Walker that many types of organic coating lose up to 85% of their initial adhesion under water-soaked conditions [63], and that adhesives show a marked loss of bond strength in water [64,65].

To improve the initial bond strength between adhesive and substrate, adhesion promoters may be used.

In order to achieve hydrolytic stability, the use of coupling agents capable of forming primary chemical bonds, (50 to 250 kcal/mol) is required [63-67].

Adhesion promoters or coupling agents, by definition, are a group of specialty bifunctional compounds that can react chemically with both the adherend (or substrate) and the adhesive [64]. The adhesion promoter forms covalent bonds, which are both strong and durable, across the interface. Adhesion promoters usually consist of molecules with short organic chains having different chemical composition on either end of the chain. One end is an organofunctional group that particularly compatible with the given adhesive material. At the other end of the chain is an inorganic functionality that is especially compatible with a given substrate. The adhesion promoter, therefore, acts as a chemical bridge between the adhesive and the substrate. As seen from Figure 2.9, an

25

adhesion promoter forms a layer between the substrate and the adhesive [64]. Usually, chemical bonds are formed between the adhesion promoter and the adhesive, and between the adhesion promoter and the substrate surface.

Adhesion promoters

Interphase region affected by adhesion promoters

-----'

~

-----,,.--

Adherends

Adhesive

Figure 2.9: Adhesion promoters provide a strong interphase region (interphase region is not drawn to scale).

Bell et al. [22] described that a polymer-substrate bond is comprised of several regions or components, as shown in Figure 2.10. The regions, and the interphases between regions, have their own unique set of properties and characteristics. When the bond is subjected to a tensile stress, the regions and interphases may be considered as links in a chain; failure will occur at the weakest linle If the weakest link is within the polymer matrix, then the improvement of adhesive strength by coupling agent will be of no benefit.

Polymer Matrix

~

Polymer/Coupling Agent

~

Interphase

\J'-~'1"~ ...

Coupling Agent Layer

,..=-<---,"---,-~~,--,----,,",--,---,-j=-, ~

Coupling Agent/Substrate

Interphase

Substrate

Figure 2.10: Schematic diagram of the polymer-coupling agent-substrate.

26

Adhesion promoters can be applied directly to the substrate in the form of a solution in a suitable solvent or solvent mixture or they can be mixed with the adhesive itself [68]. When mixed with the adhesive, the coupling agent is capable of migrating to the interface and reacting with the substrate. Only 0.1 to 2.0% by weight of adhesion promoter is added to polymer matrix prior to application. The advantage of this method is that it does not require a separate substrate coating step. The main disadvantage is that there could be stability problems in storage due to hydrolysis and polymerization. When applied directly to the substrate, adhesion promoters are applied in very thin coating that ideally is a monolayer thick. The new layer is very thin so that it provides improved interfacial bonding characteristics, yet it is not thick enough so that its bulk properties significantly affect the overall properties of the bond. More recently, experimental work has been performed by Cave and Kinloch [69] who synthesized and studied the adhesion of organosilanes with long alkyl chains, which were adsorbed on an aluminum-oxide substrate from solution. They found that, if the hydrocarbon chain (the group R') was greater than about eighteen carbon atoms (approximately 3.0 nm) in length, then the silane molecules would form a self-assembling monolayer structure. Such oriented monolayers were found to exhibit good adhesion, and impart good resistance to debonding of the interface in the presence of water.

2.3.1 Effect of the Coupling Agent Layer Thickness

A coupling agent layer thickness is one of the important factors for adhesive strength. Schmidt and Bell [70] have reported an optimum coupling agent layer thickness for polymer mercaptoester containing coupling agent bonded to an epoxy-polyamide

27

coating in the range of 150

A, several orders of magnitude less. They found failure to be near the coupling agent-polymer interface. The thickness effect data appear to be consistent with the results of Sung et al. [71], who investigated the effect of the concentration of y-aminopropyl silane (y-APS) solutions on the 180 0 peel strength of polyethylene bonded to y-APS treated Ah03.

They also found that the peel strength went through a maximum as coupling agent concentration increased.

Similarly, Plueddemann has reported [72] an optimum silane coupling agent layer thickness of 50-200

A for commercial glass fiber treatment for use in composite materials. On the contrary, optimum adhesion of polypropylene to aluminum was obtained with a silane primer layer between 0.5 and 10 JLIT1 thick. In both cases, significant deviations from the optimum thickness resulted in. substantial performance declines. Trivosono et al. [73] discovered a similar thickness dependence when silane coupling agents in fiberglass-reinforced polyester systems.

The performance of coatings and adhesives depends to a large extent on the nature of the interactions occurring at the interface between the adhesive or coating and the substrate material. These interactions may be physical, as in the mechanical interlocking that occurs when a liquid wets a rough surface and then solidifies. They may also be chemical, resulting, for example, from acid-base or van der Waals attractions.

Owing to the is the short-range nature of intermolecular interaction, this region of chemical interaction may extend no more than a few molecular layers into either phase which is referred to as the interphase region [74] to distinguish it from either the bulk phase or interface. When applied to adherend surfaces from dilute solutions, adhesion

28

promoter layers, even if only a few molecular layers thick, contain functionality much in excess of that necessary to chemically bond to the adhesive. Therefore, a thicker adhesion promoter layer does not necessarily result in improvement of hydrothermal stability of adhesive bonds. Typically, thin films only a few molecular layers thick give the best improvements in performance. They can function to improve the mechanical strength of an interphase by forming interpenetrating polymer networks with the organic phase

[75,76]. Primary chemical bonds may also be formed with the organic phase if the adhesive and adhesion promoter are chemically compatible.

2.3.2 Mechanism of Adhesion Promotion

Many theories about mechanism of adhesion promoters have been described in detail by Rosen [77] with sole reference to silanes, but it is likely that many of the proposed theories apply equally well to coupling agents in general. The proposed mechanisms are described below.

2.3.2.1 Chemical Bond Theory

As applied to silanes, the theory postulates that trialkoxysilane groups chemically bond to silinols on the mineral substrate surface by reaction of hydrolyzed alkoxy group forming interfacial bonds of 50-100 kcallmol [78] to 50-250 kcallmol [79]. The organofunctional groups of the silane bond chemically to the polymer molecules. Both reactions were considered to be essential if true coupling is to be achieved. Although originally postulated for silanes, it is equally applicable to other adhesion promoters.

It

29

has been pointed out that although covalent bonds may be fonned between polYmer and mineral surfaces, some covalent oxane bonds are easily hydrolyzed [78,80], and examples of hydrolytically stable bonding has been achieved in the absence of chemical reaction with the polYmer [78,81].

2.3.2.2 Defonnable Layer Theory

This theory postulates that the interfacial zone is plastic, allowing stresses between the polYmer and mineral surfaces to self-relieve without bond rupture. Internal stresses are thus reduced.

It has been suggested that in the case of a silane, the film is too thin to allow this [77]. However, it is possible that the presence of the coupling agent might cause preferential adsorption. Credence for this view is given by the number of workers who report that the amount of adhesion promoter used is critical and that excessive usage may result in adhesion failure [82,83].

2.3.2.3 Surface Wettability Theory

Erickson and Plueddemann suggest that particularly in filled systems, complete wetting of the mineral surface will improve adhesion by physical adsorption that would exceed the cohesive strength of the polYmer. However, it is difficult to see how physical adsorption provides bond reinforcement when the polYmer is in competition with water and possibly other weakly bonded surface layers, and where chemical bonding is also present [78].

30

2.3.2.4 Restrained Layer Theory

In effect, this theory postulates a chemical reaction between promoter, polymer, and mineral substrate as in the chemical bond theory but also suggests that the presence of a region of intermediate modulus between polymer and substrate which transfers stress from the high modulus surface to the relatively low modulus polymer. Adhesive technology has long recognized this principle in specially formulated primers for use when bonding rubbery polymers to metals

2.3.2:5 Reversible Hydrolytic Bond Theory

Best regarded as a combination theory, it postulates the chemical reactions between coupling agent, substrate, and polymer of the chemical bond theory together with the rigid interface of the restrained layer theory and the plastic interface of the deformable layer theory.

It allows for stress relaxation by the reversible breaking of stressed bonds without loss of adhesion in the presence of water.

It also argues that when

Si-O or Ti-O substrate bonds are broken by intrusion of water, they may re-form with some recovery in adhesion.

It is likely that hydrogen bonding is a particularly important aspect of this theory, especially in the case of silanes [84]. Recovery of adhesion between urethane and epoxide coatings and metal substrates on drying out after water immersion has been demonstrated by Walker [85-87].

It is now generally accepted that some silane coupling agents do not need to react with the polymer chemically to provide enhanced initial and wet adhesion [82,88].

31

2.3.2.6 Oxide Reinforcement

This theory postulates that a primary mechanism by which silane coupling agents improve initial and wet adhesion is by reaction with the oxide surface on a metal to increase the cohesive strength of the oxide [89] and certainly, in the case of aluminum oxide, increase the wet strength of the oxide by inhibiting hydration of the oxide [90].

This has the effect of causing any failure to occur in the new weakest layer (i.e., the adhesive or coating). This may also explain, in part, the differences in bond strength

" achieved with different coupling agents and different metals, as it may be the nature of the oxide film and the degree of reinforcement that varies rather than any intrinsic property of a particular metal adhesion promoter combination.

It .

seems unlikely that any single theory can explain the mechanism of adhesion promotion for such diverse systems as particulate- and fiber-filled composites, surface coating, and adhesives applied to the complete range of metallic and other mineral substrates. Plueddemann suggests that all theories of adhesion describe factors that are involved in bonding through silane coupling agents [82], and this view is likely to apply generally to the entire field of adhesion promotion.

2.3.2.7 Other Mechanisms

In addition to the stated theories of adhesion promotion, there are other mechanisms that may be both operative and important and are of general application.

32

· A. Interpenetration Network Fonnation Theory

It has been suggested that in the case of the silane coupling agents, interdiffusion of siloxinol segments with polymer molecules may be a factor in bonding thermoplastic matrices. Interpenetrating polymer networks (IPNs) need not necessarily involve crosslinking of the silane or other coupling agent and the polymer matrix. The polysiloxane surface layer has an open porous structure and the liquid adhesive penetrates this and then hardens to form an interpenetrating interphase region.

It is possible that the solvent used to dilute and apply .the silane adhesion promoter opens the molecular structure on the substrate surface, allowing the silane to penetrate and diffuse into the adherend.

Pluddemann expresses the view that to establish a strong interpenetrating layers involves a tricky interplay of mechanical and chemical interreaction at the interface.

B. Acid-Base Reactions

A comprehensive account of acid-base reactions is covered in detail in volumes 4

(No.4), 5, 8 (1990) and volume 5 (No.1) (1991) of the Journal ofAdhesion Science and

Technology, and a detailed account is beyond the scope of this chapter. Since different metal oxides have different isoelectric points in water and may therefore be regarded as acidic or basic, addition of materials having acidic properties to adhesives to be used on basic substrates, or basic materials for use on acidic substrates, may improve adhesion

[90]. Work with epoxide and polyurethane coatings of similar composition to adhesives have shown them to be basic in nature; the pH of water-soaked. surfaces from which they were stripped is known to be 8 to 10. The aminosilanes APES and AAMS are strongly basic and when applied to oxide surfaces having isoelectric points in the range 9.1 (At3+)

33

to 12.0

(Fe

2

+) may be expected to produce a basic surface. Neither silane could therefore be expected to enhance the adhesion of a basic polymer by an acid-base reaction, although both have been shown to improve the initial and wet adhesion of epoxides and urethanes. This is not to argue that acid-base reactions are unimportant in adhesion promotion technology.

2.3.3 Types of Adhesion Promoters

Classifying adhesion promoters is difficult. Usually, these materials are grouped by their· chemical composition. There are several of them that are widely known nowadays including silanes, titanates, zirconates, chromium-containing promoters, phosphorus-containing promoters, amine promoters, and polymeric adhesion promoters.

The Table 2.1 shows an application of some adhesion promoters with several substrates.

Table 2.1: Plot of primary bond formation for various types of coupling agents.

Bond

Si-O-Si

Cr-O-AI

Zr-O-AI

Si-O-Fe

Si-O-Si

Si-O-Pb

Ti-O-Si

Coupling agent/substrate

Silane/Ah03

Volan/Aluminum

Zirconate/aluminum

Silane/iron

Silane/Sh03

Silane/lead oxide

Titanate/Si0

2

34

2.3.3.1 Silanes

Silanes are the most common type of commercial adhesion promoter. They are commonly used to enhance adhesion between polymeric and inorganic materials. They usually have the form X

3

Si-R, where X is typically a chlorine or alkoxy group (-OR') and

R is the organofunctionality. The organofunctional portion bonds with the resin in the adhesive or the organic medium, and the silane portion bonds to inorganic or substrate surface. Silane adhesion promoters are commonly used between the adhesive and the adherend, between resin matrix and reinforcing fibers in composites, and between resin matrix and mineral fillers in plastic compounds. The resulting interface provides [68]:

• A chemical bridge between the surface and organic polymer or between organic polymers

• A barrier to prevent moisture penetration to the interface

• Transfer of stress from the resin to the substrate or inorganic filler component thereby improving joint strength or bulk properties

• Effective dispersion of fillers and reduction in the apparent viscosity of the system

These chemicals are usually applied to fibrous reinforcements or to the substrate surface as an aqueous solution. The solutions usually are very dilute, only 0.01 to 2% by weight of silane to keep the highly reactive hydrolyzed molecules from reacting with one another. The adhesive bond strength enhancement increases with silane concentration up to a maximum of about 2%, and then the enhancement falls-off with additional concentration. Silane adhesion promoters react with water in aqueous solutions to form

35

hydrolyzed silanes, which react with the surface of the inorganic substrate. The bound silanes then polymerize, build ug layers outward from the substrate with the organic

' functionality oriented toward the adhesive. This process is shown in Figure 2.11.

R

I

R'o-Si-OR'

I

OR'

3HzO

--.~

R

I

Ho- Si-OH --.~

I

OH

Hydrolyzed Silane

R

I

3 Ho-Si-OH

I

OH

OH OH OH

~~

Inorganic Surface

/

R R

I I

R

I

0 Si-Q-Si-o-Si-O

I I f-?

H-O

~-~

I

O-H O-H

O-H

~

Hydrogen Bond Formation

R R

I I

R

I

0 Si-Q-Si-o-Si-O

I I

0 0 0

I

~

Polymerized Silane on Surface

+ 3 H

2

0

Figure 2.11: Reaction of silane adhesion promoters on substrate.

36

Hydrolysis may take place on the surface by reaction with surface water or in solution prior to application, occurring rapidly in neutral or slightly acidic water solution and only slowly in hydrocarbon solvents [91]. Aminosilanes are autocatalytic and do not depend entirely on hydrolysis for aqueous solubility [92]. Polymerization may occur not only on the surface as the silane triols, RSi(OH)3, condense to form oligomeric siloxanes via disiloxanols and trisiloxanols but also in solution before application. The seed at which this occurs and the oligomers became insoluble depends on silane concentration, solution pH, the presence of solution catalytic salts [82], and the type of silane [93].

The hydrolyzed silonol group will react with inorganic surface hydroxy groups to form hydrogen bonds, followed by condensation to form oxane bonds.

It should be note that both hydrogen and oxane bond formation is reversible.

The reactions of silane coupling agent with the polymer, in this case aminosilane with epoxy resin, is shown in Figure 2.12 [94].

UV'HC-CH -NH-R I

2

OH

Figure 2.12: Reaction of silane with polymer.

The range of silanes available commercially is large and continually expanding.

Typical silane adhesion promoters are shown in Table 2.2.

37

Table 2.2: Typical silane adhesion promoters commercially available.

Chemical description

3-Chloropropyltrimethyl CICH2CH2CH2Si(OCH3)3 oxysilane

Structure

Functional group

With polymer With substrate

ChIaro Methoxy

Vinyltriethoxysilane y-Glycidoxypropyl trimethoxysilane y-Mercaptopropyl trimethoxysilane

CH2=CHSi(OC2Hs)3 y-Methylacryloxypropyl CH2=C-C-OCH2CH2CH2Si(OCH3)3

I \I trimethoxysilane

H3C 0

Methacrylate

CH2CHCH20CH2CH2CH2(OCH3)3

V

HSCH2CH2CH2Si(OCH3)3

Vinyl

Aliphatic epoxide

Mercapto

Ethoxy

Methoxy

Methoxy

Methoxy y-Aminopropyltriethoxy NH2CH2CH2CH2Si(OC2Hs)3 silane

Amino Ethoxy

N-~-(Aminoethyl)

Aminopropyl trimethoxysilane

NHzCH2CH2NHCH2CHzCHzSi(OCH3)3

Aminodiamino Methoxy

2-(3,4-epoxycyclohexyl trimethoxysilane

°OCH2CH2Si(OCH,), Cycloaliphatic Methoxy epoxide

38

From the Table 2.2, there are a number of silane adhesion promoters available, and they differ from each ather in the degree of their reactivity. Silanes amy be produced with amino, vinyl, epoxide, methacrylate, mercapto, and other functionalities. Each type of silane works well with different types of polymer. Table 2.3 shows the recommended silanes for various resins.

Table 2.3: Recommended silane adhesion promoters for various resins.

Silane functionality

Vinyl

Applications

Free radical cure systems: crosslinked polyethylene, peroxide cured elastomers, polyesters.

Polyethylene.

Polypr-opylene

Epoxide

Methacrylate

Epoxy, acrylics, urethanes, polysulfide

Unsaturated polyester, acrylic

Amino

Mercapto

Nylon, urethane, epoxy, phenolic, butyl rubber

Sulfur cure rubber, urethane, epoxy, polysulfide

Virtually all glass fibers used in fiber glass composites are treated with silane to improve physical properties and the resistance of the composite to deterioration by water immersion. Silanes are also generally effective in improving adhesion to metals, including aluminum, steel, cadmium, copper, and zinc. Table 2.4 shows the relative influence of the type of substrate on the effectiveness of the silane adhesion promoter in

39

improving adhesion.

It should be note that smooth, high energy substrates are excellent substrates for silane attachment. Rough, discontinuous substrates show very little benefit.

Table2.4: Effect of substrate on silane adhesion effectiveness.

-=

G)

G) u

>l

~

=

Silica

Quartz

Glass

Aluminum r=

~

; Gl lI':l

III

Gl

~ u

~

~

"'Cl

CI

0 Gl

= C> i Copper

Alumina

Inorganics

Alumino-Silicates (Clays)

11ica i

Talc

Inorganic Oxides

Steel, Iron

Asbestos in Nickel

:.= t'n

Zinc

Lead i

Chalk (Calcium Carbonate)

Gypsum (Calcium Sulfate) c

Z

Barytes (Barium Sulfate)

Gl

= Graphite

Carbon Black

2.3.3.2 Titanates

Titanate adhesion promoters have a basic R'O-Ti-(OR)3 structure. The R'O is a hydrolysable moiety that attaches to the inorganic substrate as shown in Figure 2.13

Cassidy and Yoger [95] speculate that ester linkages are hydrolyzed and coordination or

40

condensation occurs between the resulting hydroxyl groups and substrate surface groups.

However, there is no condensation polymerization to produce a polymer network at the interface. Calyert an~L~cg~~~J)!ker

[96] infer the presence of strong bonds between isopropoxytitanium tristearate and Si0

2 and Ah03 by the failure to remove the coupling agent by an extended hot-water treatment.

OR

OH

+ x R'O-Ti-(OR)3 0-Ti-(OR)J

+ x R'OH

0-Ti-(OR)3

Figure 2.13: Titanate condensation on a hydroxyl containing surface.

A comprehensive patented range of titanates is marketed by Kenrich

Petrochemical, Inc. under the trade name Ken-React, and most of the published data on titanates emanates from this source. Titanate adhesion promoters allow increased impact strength and other physical properties in composites and highly filled plastic compounds.

Titanate adhesion promoters can provide a dual function of improving adhesion and processibility. Titanates have been used predominantly to modify the viscosity of filled systems [97].

It has been shown that a small percentage of titanate in a heavily filled resin system can reduce the viscosity significantly. Thus, titanate adhesion promoters allow

41

higher fillings of particulate matter to either improve properties or lower the cost of the system without having negative effect on the viscosity of the system.

XPS studies by Yang and co-worker [98] showed that aluminum and steel surfaces treated with di(dioctyl)pyrophosphate were covered with the titanate coupling agent, and in the case of steel, the octyl groups of the titanate molecule were uppermost, confirming the view that titanates modify hydrophilic metal oxide surface with hydrophobic organic layer. The possibility has been raised that acidic surface sites on glass may catalyze condensation with surface silanols when chelate titanium acetyl acetonate is used [99]. The range of chemical types include monoalkoxy, chelate, coordinate, neoalkoxy, and cycloheteroatom. A very few of the wide range of commercial titanate coupling agents available are shown in Table 2.5.

Q

42

Table2.5: Typical titanate coupling agent.

Chemical description Structure

CH3

I

Isopropyl tri(N-ethylamino CH3-CH-O-Ti( -O-C2~NHC2HJm2)3 ethylamino)titanate

Isopropyl triisostearoyl titanate

CH3

I

CH3-CH-O-Ti(-O-C17H3S)3

Titanium di(dioctylpyro phosphate)oxyacetate

Type

Monoalkoxy

Monoalkoxy

~

I

.

C-O" [

~ ~

.

/Ti (O-r-O-p(OCsHd2

J

Monoalkoxy

CH2"° OH

2

Tetraisopropyl di(dioctyl phosphito)titanate

(CH,-CH-O-).Ti [<HlOC,H

17 h

J

Coordinate

Neoalkoxytri [P-N-(~-amino RO-Ti(OC6~NHC2~2)3 ethyl)aminophenyl]titanate

Neoalkoxy

The trade literature [83] is emphatic that it is critical to used the correct amount of titanate coupling agent. The used of excessive amounts is probably the most significant factor in application failure tests.

It is strongly recommended that selected titanates should be examined in a range of concentrations from 0.1 to 2.0% by weight in a filled system and even lower for unfilled systems. Excess titanate will result in unreacted

43

alkoxy groups on the surface and in a loss of adhesion of the polymer. This could lead to the mistaken conclusion that a particular titanate was unsuitable or even harmful.

In general, titanates with the more polar organic moisties such as isopropyl tri(Nethylenediamino)ethyl titanate and neoalkoxy tri[P-N-(b-aminoethyl)amino phenyl]titanate are recommended for adhesion promotion to polar substrates. Titanates with relatively nonpolar moieties, such as aliphatic carboxy titanates and isopropyl tri(dioctylphosphato)titanate, will adhere better to the nonpolar substrates.

2.3.3.3 Organic Resin Adhesion Promoters

A wide variety of organic resins have been claimed to act as adhesion promoters on many different substrates. Mahajan and Ghatge have reported that the use of liquid epoxide resin improved the adhesion of polysulfide sealants to anodized aluminum alloy

[100].

Abietate-tenninated polysulfide polymers, epoxide-tenninated polysulfide polymers, and abietate-terminated polyesters have also been claimed to improve the adhesion of polysulfide sealants [101]. Oxygenated fluorocarbon adhesion promoters are claimed to improve the bonding of polyacetal and polyamide-imide substrates [102]. An adhesion promoter based on ethylene-vinyl mercaptoacetate copolymer has been shown to improve the adhesion of epoxide adhesives to steel [103]. Adhesion promoters containing diorganopolysiloxanes are claimed to improve the adhesion to silicone elastomers [104], and aromatic polyether resins with aminophosphonic acid groups are state to improve the adhesive bonding of steel and aluminum [105].

44

.2.3.4 Previous Work of Epoxy/Glass Stability

Narsavage-Heald and Pearson [106] have reported the effect of adhesion promoters on the wet adhesive strength of epoxy to glass. The effect of several silane and titanate adhesion promoters on the wet adhesive strength of an imidazole-cured bisphenol

F epoxy resin on a borosilicate glass substrate was studied. The adhesion promoters used were y-aminopropyltriethoxysilane (yAPS), aminoethylaminopropyltrimethoxysilane

(AEAP), glycidoxypropyltrimethoxysilane, 2-(3,4-epoxycyclohexyl)ethyltrimethoxy silane (ECH), neopentyl(diallyl)oxytris(m-aminophenyl)titanate (LICA 97), neopentyl(diallyl) oxy tris(N-ethylenediamino)ethyltitanate (LICA 44), and two multicomponent adhesion promoters, y-APS/phenyltrimethoxysilane (PH) and AEAP/PH were used. Each of the adhesion promoters was added to the epoxy resin in 0.2% by wt.

concentration. In the multicomponent samples, each adhesion promoter was added to the resin 0.1 % by wt. to give a total of 0.2% by wt. of adhesion promoter in the resin. The specimens were aged in a controlled humidity chamber at 85% RH/85°C for 4 days. In double cantilever beam testing, the specimens were loaded to a maximum of 250 N to test for failure. Using adhesion promoters as a surface treatment was generally superior to adding them into the resin. Glass transition temps. (T g) were typically lower than the T g of the resin itself, but in most cases the difference was insignificant. The resin containing the titanate LICA 97, however, has a T g

12° lower than the control. Addition of adhesion promoters to bisphenol F/imidazole cured epoxy in 0.2% by wt. concentration improved the wet adhesive strength of epoxy glass with y-APS providing the best performance.

45

2.3.5 Effects Other than Adhesion Promotion

It should be noted that side effects are possible when using adhesion promoters/coupling agents as additives in adhesives and coating, usually beneficial but not invariably so. Beneficial effects from the use of titanates include deagglomeration, improved wetting, improved corrosion resistance, increased hydrophobicity, electrical conductivity in conducting systems, and acid resistance. The overall rheology of filled systems may be changed by both titanates and silanes requiring products to be reformulated. Silanes, particularly amino silanes, may function as curing agents or accelerators in epoxide and urethane adhesives, thereby reducing the pot life of a mixed system. These side effects should be considered in the selection or rejection of an adhesion promoter even if adhesion is the primary concern.

2.4 Objectives

First, from the work of Narsavage-Heald and Pearson mentioned above, y-APS, amine functional adhesion promoter, was the most effective coupling agent which was showed the highest G 1e value. For ECH, epoxy functional adhesion promoter, some improvement was reported. Therefore, the objective of this study is to compare the rate of

G

1e drop of these two silane coupling agents which contain different types of functionality. Moreover, we will attempt to deepen our understanding of the loss of adhesive strength in terms of moisture uptake.

46

Secondly, the propose of this study is to investigate the wet adhesive strength in term of G le of commercial polymeric adhesion promoters from the Dow Chemical

Company under hygrothermal aging condition of 85% RH/85°C. Epoxy/glass interfaces will be used to demonstrate the effects.

47

CHAPTER 3

EXPERIMENTAL APPROACH

3.1 Materials

3.1.1 Epoxy Model: Bisphenol F Resin with 2,4.EMI Curing Agent

In this model system, the epoxy was diglycidyl ether of bisphenol F epoxy resin provided by Shell

(EPO~

Resin 862.) The molecular weight of this resin is 310 glmo!.

The curing agent was 2-ethyl-4-methyl-imidazole (2,4-EMI) from Air Products

(IMICURE® EMI-24.) The molecular weight of the curing agent is 110.2

glmo!.

The chemical structures of diglycidyl ether ofbisphenol F and 2,4-EMI are shown in Figure

3.1 and 3.2, respectively.

Figure 3.1: Diglycidyl ether ofbisphenol F.

Figure 3.2: 2-ethyl-4-methyl-imidazole.

The epoxy resin and hardener were mixed for 15 minutes in the ratio of 100:4 by weight.

Then the mixtur.e· was degassed for about 15 minutes and stored at - 45°C.

48

2,4-EMI is a substituted imidazole that can be used as a curing agent for epoxy to obtain a long pot life. The reaction between an imidazole and epoxide group involves a catalytic mechanism in which only small amounts of imidazole are needed to complete a crosslinking reaction. Generally, 0.025 to 0.05 mol of imidazole/equivalent of epoxy is used. Farkas and Strohm [107] have investigated the mechanism of cure of 2,4-

EMI/epoxy system and found that the imidazole becomes permanently attached to the polymer chain. The imidazole is an effective curing agent, operating through both the secondary amino hydrogen and the tertiary amine, in a catalytic mechanism. The sequence of reaction has been suggested as in Figure 3.3 (a).

An ~ctive hydrogen in the imidazole molecule can react with an epoxide group and thereby become incorporated into the resin. An NMR study [108] revealed that the second molecule of epoxy resin added to the second nitrogen and not to the hydroxyl group. Furthermore, an alkoxide ion can initiate further reaction with the epoxide group to create the new alkoxide ion. Under appropriate condition, etherification can occur as shown in Figure 3.3

(b).

49

50

etc.

Network Resin

(b)

+

OH

I

O-CH-CH-

I

2

-CH-CH

2-

Epoxide Homopolymerization

Figure 3.3: A reaction mechanism sequence of2,4-EMI cured Bisphenol F.

51

3.1.2 Adhesion Promoters

3.1.2.1 Monomeric Adhesion Promoters

A coupling agent, sometime called an adhesion promoter, can be used to improve the hot/wet adhesive bond strength between similar or dissimilar materials. There is a large variety of silane coupling agents with a vast range of end group functionalities such as vinyl, epoxy, acrylate, amine, mercaptan, etc. Two different types of silane adhesion promoters were employed in this study and are shown below in Figure 3.4 and 3.5. The first one is y-aminopropyltriethoxysilane (y-APS) which has an amino functional group to react with epoxy matrix. The y-APS solution (Catalog number: 44,014-0) was obtained from Aldrich Chemical Company.

The latter is 2-(3,4-epoxycyc1ohexyl)ethyltrimethoxysilane (ECH) which is composed of epoxy active end group.

Figure 3.4: y-Aminopropyltriethoxysilane.

Figure 3.5: 2-(3,4-Epoxycyclohexyl)ethyltrimethoxysilane.

52

The primary structure difference between these two molecules is the replacement of amino active end group in y-APS with epoxide end group in order to give the ECH structure. The difference of the end groups in these molecules provided for possible comparison in activity of reactions between amino and epoxide group with epoxy resin which lead to different adhesive strength at interfaces because the chemical end group react differently with monomer or polymer in contact with it.

In literature as described in

Chapter 2, it was. claimed that the organosilane that has epoxy end group gives the highest adhesion for epoxy polymer [109]. Moreover, Luo and Wong [110] studied the adhesion of epoxy underfill on benzocyclobutene passivated silicon die and polyimide passivated silicon die.

It has been found that epoxide terminiated adhesion promoter added into an epoxy underfill material showed higher die shear strength than amino terminated adhesion promoter. However Narsavage-Heald and Pearson showed that y-

APS is the best adhesion promoter for borosilicate glass/epoxy interfaces [106].

3.1.2.2 Polymeric Adhesion Promoters

Two polymeric adhesion promoters, polyhydroxyaminoethers (pHAE) and the

BLOX®200 series (BLOX) from Dow Chemical Company were employed in this study.

Both PHAE and BLOX are epoxy based thermoplastic adhesion promoters. These amorphous materials exhibit excellent adhesion to variety of substrates, good gas barrier, superior mechanical toughness and stiffness, and high quality optical clarify and color.

For the PHAE resins, maleic anhydride (MAR) was grafted onto the PHAE molecules.

The BLOX®200 Series Adhesive Resins exhibit excellent adhesion to metal, glass, polar

53

polymers like nylon and epoxy. The BLOX®200 Series Adhesive Resins can be placed in solution by using polar solvents such as dimethylformamide (DMF), tetrahydrofuran

(THF) with the addition of 1-3% water, Dowanol® PM (propylene glycol metyl ether), or by using a low molecular weight organic acid solution, such as acetic acid. The solution ofPHAE and BLOX in different solvent systems, as seen in Table 3.1 and 3.2, were used to study the effect of solvent systems on the wet adhesive strength.. All polymeric adhesion promoters were-used as received from Dow Chemical Company.

Table 3.t': The PHAE epoxy based thermoplastic adhesion promoters in different solvent systems.

Adhesion Promoter

2% PHAE (RO 5MI)

2% PHAE (RR367)

Concentration

Solvent

5% Acetic acid in water

2% Acetic acid in water

2% PHAE (R40 5MI)

2% PHAE (RO 5MI)

1 % Acetic acid in water

98% Dowanol PM

54

Table 3.2: The BLOX® epoxy based thermoplastic adhesion promoters in different solvent systems.

Concentration

Adhesion Promoter

2% BLOX220 / 2% MAH

2% BLOX220 / 2% MAH

Solvent

2% Acetic acid in water

5% Acetic acid in water

2% BLOX220 / 1% MAH

-

2% BLOX220 / 1% MAH

\

2% BLOX (RO 5MI)

2% BLOX (RO 5MI)

+ 1%y-APS (silane) at 90: 10 by weight

2% Acetic acid in water

5% Acetic acid in water

2% Malic acid in water

5% Acetic acid in water

+ 95%MeOH+5% H

2

O at 90: 10 by weight

3.1.3 Borosilicate Glass

A borosilicate glass (BOROFLOAT), from Erie Scientific Co., was used in this study as a substrate for epoxy. The glass was received in the form of 0.5" x 3.0" x 0.125" slabs. X-ray photoelectron spectroscopy (XPS) was performed on the glass to verify the chemical composition of the borosilicate glass both in the bulk and at the surface which is shown below in Table 3.3. Also, the XPS spectra were shown in Appendix 1.

55

Table 3.3: Composition (%wt) of borosilacate glass.

Source of Data

Literature (bulk)

XPS (bulk)

XPS (surface)

SiOz

81.0

81.2

91.3

NazO

4.0

4.7

2.5

Ah0

3

2.0

1.9

1.7

B

Z

0

3

13.0

12.2

4.5

3.2 Sample Preparations

3.2.1 Surface Treatment of Adhesion Promoters

For use as adhesion promoters, organosilanes are typically applied by dipping the substrate in a solution of alcohol with water added [81].

In this study, silane adhesion promoters were diluted to a final concentration of 1% by weight in 95% methanol/5% water. Solution was used within several hours of preparation. Glass Substrates were dipped for 2-3 minutes in silane adhesion promoter solution, then air dried 10 minutes.

The preparation was finalized by baking the glass substrate at 100°C for 8 minutes.

For Dow adhesion promoter, the adhesion promoters were used as received. The glass plate was first baked in an oven at a constant temperature of 120 0

e

for 1 hour to make sure that glass is free of moisture. The glass was then transferred to vacuum desiccators where they were kept till use time.

The glass was soaked with the experimental adhesion promoter for 2-3 minutes and then air dried for 10 minutes. The glass plate was then placed in an oven at a constant temperature of IlO o

e

for 8 minutes.

The oven temperature was then ramped to 160°C and maintained at that temperature for 5 minutes.

56

3.2.2 Double Cantilever Beam (DCB) Specimen

The DCB testing specimen with the (y-APS coated glass)/Epoxy/(glass coated with the experimental formulation) structures were prepared. Borosilicate glass plates

(BOROFLOAT) (OS' x 3.0" x 0.125") from Erie Scientific were used.

The glass surfaces were first degreased using isopropyl alcohol and then treated by UV 10

3 for 20 minutes. The top adherend (glass plate) was immersed in 1% solution of 3aminopropyltriethoxysilane (y-APS) in 95% methanol/5% deionized water for 2-3 minutes. The treated glass was dipped into methanol bath to rinse off excess coupling agent. The glass plate was then air dried for 10 minutes and placed in a 110°C oven for 8 minutes. The bottom adherend (glass plate) was treated with the experimental adhesion promoter for 2-3 minutes. In order to create a weaker interface for crack initiation, an

Ilmm x 17 mm area on the glass was sputter coated with Au/Pd. The top and bottom glass plates are shown in Figure 3.6 (a) and (b), respectively. The bottom plate along with

242 J.lm

shims was place in a silicone mold. The mixture of bisphenol F epoxy resin and

2,4-EM! was placed on the bottom plate. The y-APScoated glass plate was placed on the top and was allowed to settle under its own weight for a few minutes before it was gently pressed against the shims to squeeze out excess resin and any trapped air. The epoxy was cured by placing the mold in an oven at 60°C for 4 hours and then post cured at 150°C for

2 hours.

57

(a) Top glass plate (y-APS treated glass) (b) Bottom glass plate

(experimental adhesion promoter treated glass)

Figure 3.6: The top and bottom glass plates of DCB specimen.

The specimen was then cooled to room temperature. Excess epoxy was removed from the edges by a grinding method. The specimen edges were then polished with 6 micron diamond. The final DCB specimen is shown below in Figure3.7.

p

~1!III!!i!II~======IIh p

Figure 3.7: Double cantilever beam (DCB) specimen.

A started crack was created by on the Au/Pd coated end of the bottom plate by using a Jewler saw. Aluminum stubs were glued to the Au/Pd side of the bottom plate using a room temperature cured adhesive (Loctite Superbonder 409). A diagram outlining the DCB specimen preparation is shown in Figure3.8.

58

Top glass plates

Borosilicate glass plates

(0.5" x 3.0" x 0.125")

I

• Wipe with isopropanol

• Treated with UV/0

3 for 20 minutes

I

+

Bottom glass plates

I

• Dip in y-APS solution for 2-3 minutes

• Rinse with methanol

• Air dry 10 minutes

• Dry at

Boac,

8 minutes

• AulPd sputter

• Dip in experimental

Adhesion promoter

• Air dry 10 minutes

• Dry at 110 a

C, 8 minutes

• Glue 4 corners with shim for thickness control

• Place bisphenol F/2,4-EMI mixture on top of the bottom plate

• Cover with y-APS treated glass

Cure epoxy at

• 60 a

C, 4 hours

• 150 a

C, 2 hours

Ir

I

Glass/epoxy/glass structure

I

Grinding and polishing

Ir

I

DCB specimen

Figure 3.8: A diagram of DeB specimen preparation.

59

3.3 Double Cantilever Beam (DCB) Testing

DCB testing was done by computer control screw driven Instron testing machine

(model 5521) in displacement control at a speed of 0.127 mrnJmin and using 500 N load cell. Tensile force was applied at the end of a specimen in a direction normal to the crack surface which is called cleavage mode or mode 1 fracture as shown above in Figure 3.9

As the load was applied, the crack propagates, leading to a load drop the specimen is, then unloaded further using the crosshead position and subsequently reloaded using crosshead position. This cycle was repeated in order to get several strain energy release rate (G

1c) values for averaging. G

1c values can be determined from the load-displacement plot shown in Figure 3.9.

40

35

30

~

"Cl

~ e

~

25

20

15

10

5

0

0

0.1

0.2

0.3

Displacement (Imn)

0.4

0.5

Figure 3.9: Typical double cantilever beam (DCB) analysis.

0.6

60

\',

\

\~ " ,

By using the maximum load of each cycle and a crack length, G

1e can be calculated from the following relation from Boa et ai.

[111]:

Gte = where, P is the applied load, a is the crack length, w is the specimen width, h is beam height, and E is the plane strain modulus of glass (62 GPa).

3.4 Evaluation of two Silane Adhesion Promoters

3.4.1 Effect of Aging Time on G

1e

The effect of aging time on G

1e value of untreated, y-APS and ECH treated glass surface was studied. The DCB specimens were prepared and aged in a Despatch

Ecosphere Series controlled humidity chamber at 85°C/85% RH. The specimens were tested after 4, 7, 11, 14 days of aging by using a computer controlled screw driven Instron testing machine. The DCB specimens were precracked before the test was performed by using razor blade. The G

1e was measured as a function of aging time.

3.4.2 Moisture Uptake of Aging Specimen

In order to explain the drop in G

1e after humidity exposure, the moisture uptake of aging sample as a function of time was measured. The DCB specimens of untreated, y-

APS and ECH treated glass surface were prepare and aged in a controlled humidity chamber at 85°C/85% RH for 14 days. The specimens were weighted every hour in the first days and then every day until 6 days. After that, the weight was measured every 2 days. Water content, M b can be calculated by using the following equation:

61

Wt-Wd

Water content (%)

=

x 100

Wd where,

Wt is the weight of the sample at a certain moment of time and

Wd is the weight of the dry sample. Finally, the curve relationship between water content and aging time was plotted.

3.5 Evaluation of Polymeric Adhesion Promoters

The effect of two epoxy based thermoplastic adhesion promoters on the wet adhesive strength of an imidazone-cured bisphenol F epoxy resin on borosilicate glass substrate was also studied. The experimental procedure used in the study was a slightly modified version of the procedure used by D. Narsavage-Heald and R. Pearson ("Effect of Adhesion Promoters on the Wet Adhesive Strength of Epoxy to Glass", Proceedings of the 24 th

Annual Meeting ofthe Adhesion Society, pp 383-385,2001).

The specimens were aged in a Despatch Ecosphere Series controlled humidity chamber at 85% RH/85°C. The Double Cantilever Beam (DCB) test was performed on the aged using a computer controlled screw driven Instron testing machine.

If a specimen did not fail when the loading force reached to 250 N, the testing was stopped and the specimen was returned to the aging chamber. The specimens were tested after 4, 7, 14,

21,28, '" days of aging (up to 42 days). In the cases where the precrack did advance, the critical interfacial strain energy release rate (G

1c) was measured. The specimens were fractured completely when critical strain energy release rate is measure. The results are analyzed in terms of days until G

1c can be measured as well as G

1c at the time the interfacial crack is propagated.

62

CHAPTER 4

RESULTS AND DISCUSSION

4.1 Evaluation of two Silane Adhesion Promoters

From previous work by Narsavage-Heald and Pearson [106], the effect of several silane and titanate adhesion promoters on the wet adhesive strength of epoxy/glass interfaces was studied. The result showed that adhesion promoters improve the wet adhesive strength of epoxy/glass interface both using adhesion promoter as a surface treatmentand as an additive in the epoxy resin as seen in Figure 4.1 and 4.2, respectively.

y-APS, amino end group silane, showed the highest wet adhesive strength. Beside, it is obvious to see that ECH, epoxide end group silane, showed lower adhesive strength than y-APS in both cases. Therefore, y-APS is a superior adhesion promoter in the system of epoxy/glass interfaces.

63

Control

0

AEAP/Ph

LICA44

25

5.8 J/m

2

% Fail

50 75 100

13.3 J/m

2

17.9 J/m

2

..cl

~

....

LICA97

= e

....

l<tl

Q,l

=

, ECH _IIIIIDD_.

~ 8.8 J/m 2

Gly jmmDIIDDIDIIlIIIIII.IIIIIIIIIIIlIIIIIIDlIIQ

12.5 J/m

2

APS/Ph

AEAP • • •

_illli23i·4~J/.n2

liiiEEl

.i"!'i1"';;'''''''';;;;!''''';;;;!;;''''!!!'!''''';;''''ii''''''''"!,,,;;,,,,,,,,,;;,,,,;;;;,,;;,,;;,,,,"';;'

249 J/m2

APS

29.8 J/m

2 m1 % Fail 4 days

• % Fail 7 days

1DI % Fail 10 days

Ii % Fail 14 days

Ii % Fail 28 days m1 % Fail 42 days

Figure 4.1: Percent failure of DCB specimens under 85°C/85% RH. aging condition using adhesion promoter as a surface treatment.

64

o

25

0/0 Fail

50 75 100

Control

ECH

• • • • • • • • • • • • • • • • 9.1J/m

2

GLY

.....

Q

Q e LICA44

.....

l:Il

CIJ

-=

<

=

AEAP

APSlPh

AEAPlPh

5.4

J/~2

17.1

J/m

2

7.7

J/m

2

6.4

J/m

2

D111IllDD.unn(DlIIllDDlIIIIIIIIlIDDIllnnq 11.8

J 1m

2

LICA97

APS

• • _ _.6.9

J/m

2 m

% Fail 4 days

• % Fail 7 days

UII % Fail 11 days

Ii % Fail 14 days

Ii % Fail 18 days

Figure 4.2: Percent failure of DCB specimens under 85°C/85% RH. aging condition using adhesion promoter as an additive in epoxy resin.

65

4.1.1 Effect of Aging Time on Strain Energy Release Rate (G

1c)

Since using adhesion promoter as a surface treatment was generally superior to adding the adhesion promoter in the resin, this study was focus on the surface treatment approach. The G

1c drop untreated, y-APS and ECH treated DCB specimens is shown in

Figure 4.3.

50.0

45.0

40.0

35.0

~ e

..

30.0

~

25.0 -

'" t.'

20.0

15.0

10.0

5.0

0.0

0

--.- Untreated glass

--IIAPS treated glass

-+ECH treated glass

50 100 150 200

Time (hours)

250 300 350

Figure 4.3: G

1c drop of untreated, y-APS and ECH treated DCB specimens under

85°C/85% RH. aging condition.

The

G\c of unaged samples is the Gl c of the samples at the room condition of about 25°C/50% RH. which were 18.1,45.0 and 21.4

J/m

2 for untreated, y-APS and ECH treated samples, respectively. As expected, the G

1c dropped with aging time under

66

85°C/85% RH. aging condition which can be explained by loss of interfacial forces.

It was believed that physical interactions such as the hydrogen bonding between the molecular chains of epoxy and glass surface could be easily disrupted by the diffusing of water. Therefore, in case the untreated DCB specimens which have only the physical interaction at the interfaces, the

G\c dropped very fast at the beginning and reached the plateau region at about 336 hours. The adhesive strength improvement can be seen in the y-APS and ECR treated specimens both before and after aging due to the introduction of chemical bonding at the interfaces. The chemical bonding, which is able to resist to the hygrothermal condition, was formed in the interphase region. Both y-APS and ECR form the oxide linkages with glass surface but at the other side of adhesion promoter molecules, they are different.

An y-APS molecule has amino functional group forming the bond with epoxy resin while ECR molecule has epoxide functional group. The amino groups of y-APS can attack the epoxide ring of Bisphenol F part in epoxy resin. The epoxide ring of ECR could be opened by acetic acid and then bonded with the epoxide ring of Bisphenol F part in epoxy resin by etherification reaction under appropriate condition, as seen in Figure

3.3 (b). Since the reaction between amino group of y-APS and epoxide group of

Bisphenol F is preferable, the y-APS treated specimens showed higher

G\c than the ECR treated specimens.

67

4.1.2 Moisture Uptake of Aging Specimen

In order to explain the G

1c drop, the moisture uptake of the specimens were evaluated. The result of water content as a function of aging time is shown in Figure 4.4.

1.40

1.20

-.

Q

:::R

1.00

....

~

~

0

U

=

=

0.80

0.60

0.40

0.20

-'lIkUntreated glass

_ APS treated glass

......- ECH treated glass

0.00

0 50 100 150 200

Time (hours)

250 300 350

Figure 4.4: Water content (%) as a function of aging time of untreated, y-APS and ECH treated DCB specimens under 85°C/85% RH. aging condition.

At first 3 hours of aging time, water content of all specimens increased rapidly and the rate of increasing was almost identical and then the rate of moisture absorption decreased. The moisture absorption rate of untreated specimens was very slow at nearly the end of the curve and almost reached the plateau at water content of 1.27%. Applying adhesion promoters on glass surface retard water absorption of the specimens leading to

68

the lower water content of y-APS and ECH treated specimens at certain aging time.

Zanni-Deffarges and Shanahan [59] have suggested the diffusion of water at the interface of the substrate/polymer interfaces. As seen from Figure 2.6, it would seen that water is entering the system by seepage close to the interface or in the interface region by a phenomenon called 'capillary diffusion'. Moreover, it is relevant to note that the surface treatment is an important factor in controlling interfacial diffusion. Clearly, the stronger interfaces would have a lower rate of moisture absorption. From this reason, y-APS treated specimens had the lowest water content.

1.40

-,-------------------------t

1.20

U

....

~ to:

~

..-.

1.00

~

....

Q

=

=

0.80 -

0.60

0.40 -

0.20

0.00

0.00

5.00

10.00

--.- Untreated glass

_ APS treated glass

_ ECH treated glass

15.00

20.00

Figure 4.5: Water content (%) as a function of the square root of aging time of untreated, y-APS and ECR treated DCB specimens under 85°C/85% RH. aging condition.

69

The result of water content as a function of the square root of aging time are shown in Figure 4.5. The untreated samples showed the water content of 0040% after exposed to 85°C/85% RH. condition for 3 hours and then the water content increasing rate changed. ECH treated samples showed the similar shape of moisture absorption curve which has the water content of 0.40% after exposed time of 3 hours and the rate of moisture gain decreased and slower than that of untreated samples. For y-APS treated samples, the moisture absorption rate also changed at exposure time of 3 hours but the water content at that point was 2.6% which is lower than both untreated and ECH treated samples.

Clearly, the shapes of the curve in Figure 4.5 indicate two-stage water absorption.

The first step was a Fickian type and took place up to 3 hours of exposure time and at longer exposure time, the water content linearly increased with the square root of time.

Two stage absorption is one of the most general sorption features of glassy polymers

[112].

It is generally accepted that the penetrant diffuses rapidly into the polymer, which is accompanied by reversible elastic swelling of the matrix. The stress that is developed is then slowly relieved by a molecular relaxation process such that the chemical potential of the sorbed water is decreased, leading to further sorption [113].

The experimental results presented so far indicate that first stage of water absorption was related to the process of saturation of epoxy matrix controlled by geometric factor of DeB specimen and type of polymer which were the same in all samples. The second was possibly osmotic pressure driven influenced by amount of water at the interfaces. For this reason, y-APS treated samples gave the lowest water

70

20.0

-.

'E

~ u

~

-

10.0

5.0

0.0

0 50 100

1.40

1.20

150 200

Time (hours)

(a)

____ G1c

--.- Water content

250 300 350

0.00

1.00

0.80

0.60

0.40

~

~

-.

=

' - '

..

~

=

0

U

~

~

=

0.20

1.00

70.0

60.0

~ sa

40.0

~ u 30.0

cS

20.0

10.0

0.0

0 50 100 150 200

Time (hours)

(b)

___ G1c

-.- Water content

0.80

- 0.60

0.40

-.

~

'-'

~

~

~

~

..

~

=

0

U

=

0.20

350

0.00

250 300

71

-.

1:

~ u

CJ

-

25.0

20.0

15.0

10.0

5.0

0.0

0 50

1.40

1.20

100 150 200

Time (hours)

(c)

__ Glc

.........- Water content

250 300

0.20

350

0.00

1.00

0.80

0.60

0.40

l.

....

co:

~

--

....

=

U

Q

=

Figure 4.6: Relation between G le drop and water content of DCB specimens under

85°C/85% RH. aging condition of (a) untreated samples, (b) y-APS treated samples and (c) ECH treated samples.

content because they had the strongest interfaces allowing less amount of water to seep through at the interface or in the interface region by capillary diffusion.

It is very interesting to relate the G le drop with moisture uptake of the specimens.

Figure 4.6 (a), (b) and (c) show the relation between G le drop and water content ofDCB specimens under 85°C/85% RH. aging condition of untreated samples, y-APS treated samples and ECH treated samples, respectively. The results show a good influence of moisture uptake on G le drop: the higher the rate of moisture uptake, the higher the rate of

72

G

1c drop. At the beginning the water content increased rapidly leading the dramatically drop of G

1c which can be explained in term of the stress developed by swelling of matrix and the loss of physical interactions. In case of untreated samples, there is no chemical bonding at the interface between epoxy and glass; therefore the

G\c was very low and reached plateau at about 5.1 J/m

2 after 336 hours of aging becausy water absorbed at the interfaces attack the physical interactions. y-APS and ECH treated samples showed higher G

1c at about 27.4 and 9.3 J/m

2

, respectively after 336 hours of aging due to the chemical bonding introduced by adhesion promoters. Bond breaking and chain pull out mechanism can be used to explain the more energy requirement for crack to propagate.

4.2 Evaluation of Polymeric Adhesion Promoters

Since y-APS showed the best adhesion strength of epoxy/glass interfaces, as shown earlier in this chapter, it was used to represent the monomeric adhesion promoter.

And the untreated sample was used as the control in order to determine the wet adhesive strength improvement of polymeric adhesion promoters. Figure 4.7 shows percent failure and G

1c ofDCB specimens of untreated and y-APS treated samples.

73

o

25

I-l

~

§

I-l

~

.....

r/)

=

Untreated

Q,)

=a

Silane (APS)

<

~:::::::::::::::134.3

J/nrmI % Fail 4 days

I1l1l % Fail 21 days

• % Fail 42 days

Iil % Fail 49 days

% Fail

50

75

100

Figure 4.7: Percent failure and Gl c of DeB specimens of untreated and y-APS treated samples.

Untreated sample completely failed within 4 days of aging and gave G lc of 6.8

J/m

2

An y-APS, one of silane adhesion promoter, showed a lot improvement in wet adhesive strength. It took 49 days to get 100% failure of the y-APS treated specimens and the G lc was 18.1

J/m

2

Since the G lc of 42 day aging was 18.9

J/m

2 which was almost identical with that of 48 day aging, the plateau seems to be reached.

74

4.2.1 The PHAE Series Polymer Adhesion Promoters

0 25

% Fail

50

16.5J/m

2

75

2% PHAEin

~

..

....

~

~ e

..

5 % Acetic Acid

2% PHAEin

2% Acetic Acid

....

~

CI.l

~

.::

'C

=

<

2% PHAEin

1% Acetic Acid

2 % PHAEin

DowanolPM

100

12.4J/m

2

13.0 J/m

2

30.6J/m

2

13.9J/m

2

8.9 J/m

2

32.9J/m

2

21.7 J/m

2

25.3 JIm

2

• % Fail 7 days lm % Fail 14 days lID % Fail 28 days iii % Fail 35 days

Figure 4.8: Percent failure and

G\c of DCB specimens treated by the PHAE polymeric adhesion promoter in different solvent systems.

Percent failure and

G\c of DCB specimens treated by the PHAE polymeric adhesion promoter in four different solvent systems are shown in Figure 4.8. As seen from the Figure 2% PHAE in Dowanol® PM, hydroxy group containing solvent, was the best system in PHAE series because the specimens can resist to crack for 28 days under

75

aging condition of 85°C/85% RH. and the highest G

1e in this set of experiment was reported. For the variation of acetic acid, the results show that using 1 % acetic acid aqueous solution as a solvent gave the highest wet adhesive strength.

Figure 4.9 (a) and (b) show the Gl e after 7 days and 14 days, respectively. The

PHAE adhesion promoters in all solvent systems showed improvement in wet adhesive strength compared to the untreated samples. However, y-APS treated specimens gave the highest wet adhesive strength in term of G

1e values.

It is easily to see that at a certain time, both at 7 and 14 days, the 2% PI:IAE in 1% acetic acid aqueous solution had the highest G

1e when compared with PHAE in 2% and 5 % acetic acid.

It is possible that excess acetic acid molecules can react with hydroxy groups on glass surface by esterification reaction, so there was less active site on glass surface that can react with the

PHAE adhesion promoter.

It is confirmed again by both Figure 4.9 (a) and (b) that

Dowanol® PM was the best solvent system for PHAE series polymeric adhesion promoter for epoxy/glass interfaces since they showed the highest G

1e among PHAE treated samples. Moreover; at 7 days of 85°C/85% RH. aging, PHAE in 1% acetic acid solution and in Dowanol® PM had G

1e values closed to that of y-APS treated samples (G

1e

=

34.1

J/m2) .

76

0

Gte

2

(JIm)

5 10 15 20 25 30 35 40 45

Untreated

~

....

0

0

~

~ a

2% PHAEin

5 % Acetic Acid

2% PHAEin

0

=

2 % Acetic Acid ...

rI.l

~

2% PHAEin

~

1 % Acetic Acid

<

2 % PHAEin

DowanolPM

Silane (APS)

16.5 JIm

2

0 5

34.3 JIm

2

10 15

(a)

G le

2

(JIm)

20 25 30 35

Untreated

...

rI.l

~

.=

"'Cl

<

~

~

~

~

....

2% PHAEin

5 % Acetic Acid

2% PHAEin

0

=

2 % Acetic Acid

2% PHAEin

1 % Acetic Acid

2 % PHAEin

DowanolPM

Silane (APS)

13.0

JIm

2

27.4

JIm

2

(b)

Figure 4.9: G le of DeB specimens treated by the PHAE series polymeric adhesion promoter in different solvent systems at aging time of (a) 7 days and (b) 14 days.

77

4.2.2 The BLOX®200 Series Polymeric Adhesion Promoters.

The percent failure and Gte of DCB specimens using theBLOX®200 series as an adhesion promoter of epoxy/glass interfaces are shown in Figu~e 4.10.

o

25

% Fail

50 75

5.4J/m

2

100

2% BLOX/2%MAH in 2% Acetic Acid

2% BLOX/2%MAH in 5% Acetic Acid

16.0J/m

2

~

~

e

2% BLOX/l%MAH

~:::=:~:~:::~1.~8~J/~m~2~ c in 2% Acetic Acid 1==

~

....

c: c

~

.c:

in 5% Acetic Acid

"0

-<

2% BLOX/l %MAH

8.8

2% BLOXin

.J

8.9

2 % Malic Acid

J/m

2

J/m

2

2% BLOXin

2% Malic Acid +

Silane iii % Fail 4 days

• % Fail 7 days

~ % Fail 14 days

III % Fail 21 days

E!II

% Fail 28 days iii % Fail 35 days

Figure 4.10: Percent failure and Gte of DCB specimens treated by the BLOX®200 series polymeric adhesion promoter in different solvent systems.

78

As seen from the Figure, 2% BLOX/2% MAR, BLOX-graft-MAH, showed higher adhesive strength than 2% BLOX/l % MAR.

It can be explained that MAR is the part that has active functional groups, so the higher MAR content provides more active sites reacting with either epoxy resin or glass surface. In short time period, the system of

2% BLOX/2% malic acid seems to give the highest stability of epoxy/glass interface under hygrothermal condition which can be considered from the higher G

1c'

Mixture of

2% BLOX/2% malic acid and y-APS improved adhesive strength due to the strong interaction of y-APS at interphase region.

The effect of acetic acid concentration in the solution was also studied in the 2%

BLOX/2% MAR and 2% BLOX/l % MAR systems. The results are easily seen in Figure

4.11 (a) and (b) which compare G

1c values at aging time of 14 and 28 days, respectively.

In Figure 4.11 (a), the G

1c of untreated samples and APS treated sample are also shown for comparison. The controversy was observed: in the BLOX/2% MAR system, 5% acetic acid aqueous solution showed higher G

1c values while in the BLOX/l % MAR, 5% acetic acid aqueous solution showed lower G

1c values.

Briefly comparison, it can be said that the BLOX®200 series has higher ability than the PHAE series to improve stability of epoxy/glass interfaces under hygrothermal conditions.

79

o

5 25 30 35

Untreated

'"' 2% BWXl2%MAH

~ in 2% Acetic Acid .,

I:

~

...

8

~

<

2% BWXl2%MAH in 5% Acetic Acid

2%BLOX/l%MAH in 2% Acetic Acid

2%BLOXll%MAH in 5% Acetic Acid

Silane (APS)

11.8

Jim

2

27.4

Jim

2

5

(a)

Gle

2

(Jim)

10

2% BLOXl2%MAH in 2% Acetic Acid

~

§

2% BLOXl2%MAH

~ in 5% Acetic Acid

~

2% BLOXI I%MAH

< in 2% Acetic Acid o

15

2% BLOXlI %MAH in 5% Acetic Acid

(b)

Note: G

1c data of untreated and y-APS treated samples at 28 days is not available.

Figure 4.11: G

1c of DeB specimens treated by the BLOX®200 series polymeric adhesion promoter in different solvent systems at aging time of (a) 7 days and (b) 14 days.

80

CHAPTERS

CONCLUSIONS

The effect of hygrothermal aging of 85°C/85% RH. on adhesive strength of epoxy/glass interfaces was studied. The rate of strain energy release rate (Gte) drop of two silane adhesion promoters, y-aminopropyltriethoxysilane (y-APS) and 2-(3,4-

Epoxycyclohexyl)ethyltrimethoxysilane (ECR), was investigated. The Gte of unaged samples were 18.1, 45.0 and 21.4 J/m

2 for untreated, y-APS and ECR treated samples, respectively. As expected, the Gte dropped with aging time which can be explained by loss of interfacial forces. The adhesive strength improvement can be seen in the y-APS and ECR treated specimens both before and after aging due to the introduction of chemical bonding at the interfaces. The chemical bonding, which is able to resist to the hygrothermal condition, was formed in the interphase region.

Moreover, the loss of adhesive strength was rationalized in terms of moisture uptake. The results showed that applying adhesion promoters on our glass surface retards water absorption in the DCB specimens leading to the lower water content of y-APS and

ECR treated specimens at certain aging time. Two-stage water absorption was observed.

The first step was a Fickian type and took place up to 3 hours of exposure time and at longer exposure time, the water content linearly increased with the square root of time.

The first stage of water absorption was related to the process of saturation of epoxy matrix controlled by geometric factor of the specimen and type of polymer. The second was possibly osmotic pressure driven influenced by amount of water at the interfaces.

81

Furthermore, the wet adhesive strength in term of

Gte of commercial polymeric adhesion promoters, PHAE series and BLOX®200 series, under hygrothermal aging was investigated.

It was found that 2% PHAE in Dowanol® PM, hydroxy group containing solvent, was the best system in PHAE set. The 2% PHAE in 1% acetic acid aqueous solution had the highest Gte when compared with 2% PHAE in 2% and 5 % acetic acid.

It is possible that excess acetic acid molecules can react with hydroxy groups on glass surface by esterification reaction. For BOLX®200 series, 2% BLOXl2% MAR showed higher adhesive strength than 2% BLOXll % MAR.

It, can be explained that MAR is the part that has active functional groups, so the higher MAR means the higher active sites reacting with either epoxy resin or glass surface. In conclusion, it can be said that the

BLOX®200 series has higher ability than the PHAE series to improve stability of epoxy/glass interfaces under hygrothermal conditions; however, y-APS is still the best adhesion promoter.

82

REFERENCES

1.

Lee, H.; Neville, K., Handbook of Epoxy Resins; McGraw Hill: New York, 1967.

2.

Briuns, P. E, Ed., Epoxy Resin Technology; Interscience Publishers: New York,

1968.

3.

Nicodemo, L.; Bellucci, F.; Marcone, A.; Monetta, T., J.

Membr. Sci.

1990, 52,

393.

4.

Schultz, W. l, International Encyclopedia of Composites, Vol. 2; Lee, S. M., Ed.;

VCHPublishers: New York. 1990,77.

5.

Moy, P.; Karasz, F. E. in Water in Polymers; ACS Symposium Series 127;

American Chemical Society: Washington, DC, 1980, p.505.

6.

Ellis, T.

S.; Karasz, F. E., Polymer 1984, 25, 664.

7.

McKague, E.

L.; Reynolds, 1.

D.; Halkias, 1.

E., J.

Appl. Polym. Sci.

1978,22,394.

8.

Theocaris, P. S.; Kontou, E. A.; Papanocolaou, G. C., Colloid Polym. Sci. 1983,

261,394.

9.

Fowkes, EM.; Dwight, D. W.; Cole D. A., J.

Non-Cryst. Solids 1990, 120,47.

10.

Pantano, C.

G., Rev. Solid State Sci.

1989,3,379.

11.

Feuston, B. P. and Garofalini, S. H.

J.

Appl. Phys.

1990, 68,4830.

12.

Hunston, D.; Macturk, K.; Schultheisz, C.; Holmes, G.; McDonough, W.; Schutte,

C. L. in EURADH'96, Cambridge, 1996, Vol. 1, London: The Institute of

Materials, 1996, p. 427-432.

13.

Schutte, C. L.; McDonough, W.; Shioya, M.; McAuliffe, M.; Greenwood, M.,

Composites 1994,25,617.

83

14.

Osterholtz F. D.; Pohl,E.

R., J.

Adhesion Sci.

Technol.

1992,6, 127.

15.

Plueddemann, E. P. Fundamentals of Adhesion, Lee L. H., Ed., Plenum Press, New

York, 1991, p. 279-290.

16.

Dwight, D. W.; Fowkes, F.

M.; Cole, D. A., Kulp, M. 1.; Sabat, P. J.; Salvati, Jr. L.;

, \ "j

Huang, T. C., Acid- Baase Interactions: Relevence to Adfhesion Science and

Technology, Mittal K.

L.; Anderson, Jr. H. R., Ed., Utrecht, The Netherlands: VSP,

1991, p. 243.

17.

Plueddemann, E. P., Silane Coupling Agents, Plenum Press, New York, 1982.

18.

Miller, 1. D.; Ishida, H., Fundamentals of Adhesion, Lee.

L.

H., Ed., Plenum Press,

New York, 1991, p. 291-324.

19.

Miller, 1. D.; Ishida, H., Silane, Surfaces and Interfaces, in Chemically Modified

Surfaces, Leyden, D. E., Eq., Vol. 1, Gordon and Breach, New: !,ork, 1986, p. 525.

20.

Culler, S. Ph.D. Thesis, Case Western Reserve University, 1984.

21.

Chiang, C. H.; Koening, 1.

L., SPI 35 th

Ann Tech. Conf. Reinf. Plast., 1980:B:20.

22.

Bell, J. P.; Schmidt, R. G.; Malofsky, A.; Mancini, D., J.

Adhesion Sci.

Technol.

1991,5,927.

23.

Pearson, R. A.; Lloyd, T. B.; Azimi, H.

R.; Hsiung, J.; Early, M. S.; Brandenburger,

P. D., IEEE Trans. Comp., Packaging, Manufact. Technol., Part A, Vol. 20, 1997,

31.

24.

Lloyd, T. B., Colloids and Surfaces A:Physicochemical and Engineering Espects,

Vol 93, 1993,25.

25.

Gent, A.

N.; Kinloch, A.

1., J.

Polym.

Sci., Part A-2 1971,9,659.

84

26.

Kinloch, A. 1., Adhesion and Adhesives: Science and Technology, Chapman and

Hall, London, 1987, p. 264.

27.

Irwin, G. R; Kies, 1. A., J.

Welding 1954,33, 193.

28.

Griffith, A. A., Phil. Trans. Roy. Soc.

1921,221, 163.

29.

Ivanova, K.

1.; Pethrick, R A.; Affrossman, S., 1.

ofAppl. Polym. Sci.

2001,82,

3468.

30.

Kinloch, A. J., Adhesion and Adhesives: Science and Technology, Chapman and

Hall, London, 1987, p. 350.

31.

Wong, T.

c.,

Broutman, L. 1., Polym. Eng. Sci.

1985,25,521.

32.

Diamant, Y.; Marom, G.; Broutman, L. 1., J.Appl. Polym. Sci.

1981,26,3015.

33.

Adamson, M. 1.,1.

Mater. Sci.

1980, 15, 1736.

34.

May, P.; Karasz, F. E., Polym. Eng Sci.

1980,20,315.

35.

Alfrey, T.; Gurnee, E. F.; Lloyd, W. G., J. Polym. Sci., 1966,1,249.

36.

Browning, C. E.; Husman, G. E.; Whitney, J. M. in Composite Materials: Testing and Design, ASTM STP 617, Davis Jr., 1. G., Ed., American Society for Testing and Materials, Philadelphia, 1977, p. 481-496.

37.

Whitney, 1. M; Browning, C. E. in Advanced Composite Materials-Environmental

Effects, ASTM STP 658, Vinson, 1. R, Ed., American Society for Testing and

Materials, Philadelphia, 1978, p. 43-61.

38.

Shen, C. H.; Springer, G. S., J Compos. Mater 1976, 10, 2.

85

39.

Shirrell, C. D. in Advanced Composite Materials-Environmental Effects, ASTM

STP 658, Vinson J.

R, Ed., American Society for Testing and Materials,

Philadelphia, 1978, p. 21-42.

40.

Crossman, F.

W.; Mauri, R E.; Warren, W. J.

in Advanced Composite Materials-

Environmental Effects, ASTM STP 658, Vinson J.

R, Ed., American Society for

Testing and Materials, Philadelphia, 1978, p. 205-220.

41.

Hliger, J. 1.; Schneider, N. S., Polym. Eng. Sci.

1980,20,310.

42.

Apicella, A., International Encyclopedia of composites, Vol. 2; Lee, S. M., Ed.,

VCH Publishers, New York, 1990, p.46.

43.

Hayward, D.; Hollins, E., Johncock, P.; McEwan, 1; Pethrick, R.

A.; Pollock, E.

A,

Polymer 1997, 38, 1151.

44.

Cotinaud, M.; Bonniau, P.; Bunsell, A R, 1.

Mater. Sci.

1982,17,867.

45.

Apicella, A. Estiziano, 1.; Nicolais, 1.; Tucci, V., J.

Mater. Sci.

1988,23, 729.

46.

Jacobs, P. M.; Jones, F.

R, 1.

Mater. Sci.

1989,24,2331.

47.

Jacobs, P. M.; Jones, F.

R, J.

Mater. Sci.

1989,24,2343.

48.

Suziki, T.; Oki, Y.; Numajiri, M.; Miura, T.; Kondo, K; Shiomi, Y.; Ito, Y.,

Polymer 1994, 37, 3025.

49.

Vrentas, J.

S.; Duda, J.

1., Encyclopedia of Polymer Science and Engineering, 2 nd ed., Vol. 5; Wiley, New York, 1986, p.36.

50.

Berens, A R; Hopfenberg, H.

B., Polymer 1978, 19,489.

51.

Xiao, G. Z.; Shanahan, M. E.

R., J.

Polym. Sci., Polym. Phys Ed 1997, 35, 2659.

52.

Vrentas, J. S.; Vrentas, C.

M., J;Polym. Sci., Polym. Phys. Ed.

1992,30, 1005.

86

53.

Illinger,1. L.; Scheider, N. S. in Water in Polymers; ACS Symposium Series 127;

American Chemical Society: Washington, DC, 1980, ref. 1, p. 571.

54.

Bonniau, P.; Bunsell, A. R., J. Compos. Mater.

1981,15,272.

55.

Adamson, M. J., 1.

Mater. Sci., Polym. Phys. Ed.

1997,35,2659.

56.

Yang, F.; Gilbert, R. D.; Fornes, R. E.; Memory, J. D., J. Polym. Sci., polym. Chem.

Ed.

1986,24,2609.

57.

Carfagna, C.; Apicella, A.; Nicolais, L., J. Appl. Polym. Sci.

1982,27, 105.

58.

Ivanova, K.

I.; Pethrick, R. A.; Affrossman, S., J. Appl. Polym. Sci.

2001, 82, 3477.

59.

Zanni-Deffarges, M. P.; Shanahan, M. E. R., Int. J. Adhesion and Adhesives 1995,

15, 137.

60.

Xiao, G. Z.; Delamar, M.; Shanahan, M. E. R., J. Appl. Polym. Sci.

1997,69,449.

61.

Xiao, G. Z.; Shanahan, M. E. R., J. Appl. Polym. Sci.

1998,69,363.

62.

Xiao, G. Z.; Shanahan, M. E. R., Polymer.

1998, 39, 3253.

63.

Walker, P., Olfic. Dihest 1965, 37, 1561.

64.

Falconer, D. J.; McDonald, N.; Walker, P., Chem. Ind., July 1964, p. 1230.

65.

Gledhill, R.A.; Shaw, S. 1.; Tod, D.

A., Int.

J.

Adhesion andAdhesiyes 1990,10,

192.

66.

Bjorksten, J.; Yaeger, L. L., Mod. Plastics 1952,29, 124.

67.

Bradley, D. C.; Hancock, D. C.; Wardlaw, W., J.Chem. Soc.

1952,2773.

68.

Petrie, E. M., Hand book of Adhesives and Sealants, McGraw-Hill, New York,

2000, p.

253.

69.

Cave, N. G.; Kinloch, A.

J., Polymer 1992,33, 1162.

87

70.

Schmidt, R. G.; Bell, 1. P., 1. Adhesion 1989,27, 135.

71.

Sung, N. H.; Kaul, A.; Ni, S. in Adhesion Aspects of Polymeric Coatings, Mittal, K.

L., Ed., Plenum press, New York, 1983, p. 379.

72.

Plueddemann, E. P., Silane Coupling Agents, Plenum press, New York, 1982, p 82-

85.

73.

Trivosono, N. M.; Lee, L.

H.; Shriner, S. M., Ind. Eng. Chern.

1958,50,912.

74.

Drzal, L.

T.; in Advances in Polymer Science, Dusek, K., Ed., Vol. 75, Springer-

Verlag, New Yark, 1986.

75.

Sung, N.

H.; Kaul, A.; Chin, 1.; Sung, S. P., Polym. Eng. Sci.

1982, 22, 637.

76.

Plueddemann, E. P.; Stark, G.

L., Proc. 35 th

Annu. Tech. Conf., SPI Reinforced

Plastics/Composite Inst., Sect. 20-B, 1980.

77.

Rosen, M. R., 1.

Coating Technol.

1978,50,644.

78.

Erickson, P. W.; Plueddemann, E. P., Composite Materials, Vol. 6, Academic Press,

New York, 1974, Chap. 4.

79.

Southeng, W., Polymer Interface and Adhesion, Marcel Dekker, New York, 1982.

80.

Bascom, W. D., Composite Materials, Vol. 6, Academic Press, New York, 1974.

Chap. 2.

81.

Walker, P., J.

Oil Colour Chemists' Assoc.

1982, 66, 415.

82.

Plueddemann, E. P., Silane Coupling Agents, Plenum press, New York, 1982.

83.

Technical Bulletin KR-1084-2, Kenrich Petrochemicals, Inc., Bayonne, 1987.

84.

Plueddemann, E. P., J.

Paint Technol.

1970,42, 600.

85.

Walker, P., J.

Oil Colour Chemists' Assoc.

1982, 65,436.

88

86.

Walker, P., 1.

Oil Colour Chemists' Assoc.

1983, 66, 188.

87.

Walker, P., 1.

Oil Colour Chemists' Assoc.

1984, 67, 108.

88.

Walker, P., 1.

Oil Colour Chemists' Assoc.

1984, 67, 126.

89.

Walker, P., 1.

Adhesion Sci. Technol..l991, 5, 279.

90.

Bolger, 1. C.; Michaels, A.

S., in Interface Conversion for Polymer Coatings, Weiss,

P., Cheeves, G. D., Eds., American Elsevier, New York, 1969.

91.

Plueddemann, E. P., Soc. Plast. Ind. 19A 1969.

92.

Silane Coupling Agents, Dow Coming Corp., Midland, 1981.

93.

Silane Coupling Agents, Dow Coming Corp., Midland, 1970.

94.

Walker, P., J.

Coatings Techno!.

1980,52,49.

95.

Cassidy, P. E.; Yager, B. J., Tracor Document TTO-AV-7362-U, NASA Contract

NA58-24073, 1971.

96.

Calvert, P. D.; Lalanandham, R.

R; Walton, D. R. M., in Interfacial Coupling by

Alkoxytitanium and Zirconium Tricarboxylates, Mittal, K.

L., Ed., Plenum Press,

New York, 1983.

97.

Waddington, S.; Briggs, D., Polym. Comm.

1991,32,506.

98.

Yang, C.

Q.; Moulder, J. F.; Fateley, W., 1.

Adhesion Sci. Technol.

1988,2, 11.

99.

Wang, B., Ph.D. Dissertation, University of Massachusette, Feb. 1989.

100. Majahn, S. S.; Ghatge, N. D., Int.

J.

Polym. Mater.

1983,10, 141.

101. U.S. patent 3,838,078.

102. Japanese unexamined patent 62/125,092.

103. Schmidt, R. G.; Bell, J. P., J.

Adhesion 1989, 24, 135.

89

104. European patent application 336,854.

105. U.S. patent application 4,781,984.

106. Narsavage-Heald, D.; Pearson, R.

A, Proc. Annu. Meet. Adhes. Soc. 2001, 24 th

,

383-385.

107. Farkas, A; Strohm, P.

F., J.

Appl. Polym. Sci., 1968,12.

159.

108. Dyakonov, T.; Chen, Y; Holland, K; Drbohlav, 1.; Bums, D.; Velde, D. V.; Seib,

L.; Soloski, E.

1.; Kuhn, 1.; Mann, P.

1; Stevenson, W.

T. K., Polymer Degradation

and Stability 1996,53,217.

109. Plueddemann, E.

P., in Encyclopedid of Polymer Science and Engineering, 2 nd

Ed.,

Kroschwitz, J.

I., Ed., Vol. 4. Wiley, New York, 1985.

110. Luo, S.; Wong, C. P., IEEE Trans. Comp., Packaging, Manufact. Technol., Vol. 24,

2001,38.

111. Bao, G.; Ho. S.; Suo. Z.; Fan, B., Int.

J.

Solids Struct. 1992,29, 1105.

112. Fujita,H., Adv. Polym. Sci. 1961, 71, 1.

113. Barrie, 1 A, Diffusion in Polymers; Crank, 1, Park, G. S., Eds., Academic, New

York,1964.

114. Nagarajan, K.; Dauskardt, R. H., SRC Consortium Meeting, Albany, New York,

October, 2000.

90

APPENDIX 1: X-ray photoelectron spectroscopy (XPS) analysis of borosilicate glass. .

XPS analysis of borosilicate glass was done both at the surface and bulk in order to compare the compositions. The surface results are shown in Table ALl and Table

A1.2 while the bulk results are shown in Table A1.3 and Table AlA.

Figure ALl and

Figure A1.2 show the XPS spectra of surface and bulk, respectively.

Table Al.1: XPS analysis of borosilicate glass surface.

Name Area (cps) ensitivity

) o

Is 20827

B Is 170

Na Is

Si 2p

Al 2p

1243

3926

62

Atomic Cone.

(%)

60.612

2.888

1.833

33.942

0.726

Weight

969.7234

31.2224

42.1318

953.2887

19.5913

Weight

%

48.102

1.549

2.090

47.287

0.972

Table Al.2: Composition of borosilicate glass surface.

Name

Si0

2

Na20

B

2

0

3

Ah0

3

M.W.

60.084

61.979

69.619

101.961

Molecular %

33.942

0.917

1.444

0.363

Weight

2039.3711

56.8038

1001'5298

37.0118

Weight %

91.299

2.543

4.501

1.657

91

Table Al.3: XPS analysis of bulk borosilicate glass.

Name Area (cps) Sensitivity o

Is

B Is

Na Is

Si

Al

2p

2p

2.8370

0.4860

5.6000

0.9550

0.7050

Atomic Cone.

(%)

59.762

7.429

3.252

28.747

0.810

Weight

956.1322

80.3149

74.7635

807.3882

21.8554

Weight

%

49.274

4.139

3.853

41.608

1.126

Table Al.4: XPS analysis of bulk borosilicate glass.

Name

Si0

2

Na

2

0

B

2

0

3

Ah 0

3

M.W.

60.084

61.979

69.619

101.961

Molecular %

28.747

1.626

3.715

0.405

Weight

1727.2347

100.7779

258.5998

41.2942

2127.9066

Weight %

81.171

4.736

12.153

1.941

92

\0

.j:::..

[Counts]

~

0-

0

3

;!1.

o'

......

(1l

(JQ

:n

0'Cl

....

., =

!'t>

....

en rn

"C:l

(1l

~

~

""I

C

8

50000

40000

30000

20000

10000

Na 1s

,

I

1000

I

800

J o

1s

I i

I

I

,

600

Binding Energy leV]

I

400

I

B

1~11

I

Si2s

AI2s

Si 2p3/2

Si 2p1/2

AI2p32

AI2p12

Na2 o

2s

Pass Energy

300.0

Lens Mode

Transmission

Acquisition Mod

Swept

Start Energy

1150.0000

I

1;

!I

Stop Energy

0.0000

Step Energy

1.0000

Channels

1151

Time / Step

0.300

Sweeps

1

Total Time

00:05:45

I

200 0

614101 4:02:01

(7

APPENDIX 2: Percent failure and G

1c of DCB specimens treated by polymeric adhesion promoters.

Table A2.1: Percent failure and G le of DCB specimens treated by PHAE series polymeric adhesion promoters.

4 days

Adhesion promoter Failure

(%)

7 days

Gte

Failure

(%)

Gte

14 days 21 days 28 days 35 days

Failure

(%)

Gte

Failure

(%)

G lc

Failure

(%)

Gte

Failure

(%)

Gte

2% PHAE in

5% Acetic Acid

2% PHAE in

2% Acetic Acid

2%PHAEin

1% Acetic Acid

2%PHAEin

Dowanol®PM

50

25

·50

50

16.5

18.9

30.6

32.9

100

100

75

75

12.4

13.0

13.9

21.7

100 25.3

100 8.9

Note: G le is in the units of J/m

2

95

Table A2.1: Percent failure and G le of DeB specimens treated by BLOX®200 series polymeric adhesion promoters.

4 days 7 days 14 days 21 days 28 days . 35 days

Adhesion promoter Failure

(%)

Gte

Failure

Gte

Failure

(%) (%)

Gte

Failure

(%)

Gte

Failure

(%)

Gte

Failure

(%)

Gte

2% BLOXl2%MAH in 2% Acetic Acid

2% BLOXl2%MAH in 5% Acetic Acid

2% BLOXII %MAH in 2% Acetic Acid

2% BLOXII %MAH in 5% Acetic Acid

2%BLOXin

2% Malic Acid

2%BLOXin

2% Malic Acid+Silane

25 36.2

50

75

-

16.0

19.2

25

25

75

25

25

6.6

11.0

11.8

10.8

23.6

75

50

75

100

5.4

10.5

6.9

8.9

100

100

100

100

50

3.4

9.8

8.8

2.3

U.8

.

100 15.6

Note: G le is in the units of J/m

2

96

APPENDIX 3: Study of subcritical crack growth of epoxy/glass interfaces.

Subcritical crack growth at interfaces can be characterized by applying ASTM

D3433 (Fracture Strength in Cleavage of Adhesive in Bond Joints). First, the DCB specimen, as shown in Figure A3.1, is prepared, and applied under the force to initiate the crack.

F

F

Figure A3.1: Double Cantilever Beam (DCB) specimen.

After the crack is initiated, the specimen is held at a certain extension (fix displacement). The curve relationship between load and time, as seen in Figure A3.2, is recorded. Then crack growth rate (da/dt) can be calculated by using equation []: where a is crack length at any time, t, P is the applied load, Po is the initial load, and ao is the initial crack length.

97

__ displacement fixed f

~~Iaxation

-----------

-----

Time (sec)

"

,'"

Crack Tip Driving Energy, G

Figure A3.2: (a) curve relationship between load and time; (b) curve relationship between crack growth rate and strain energy release rate (G) at crack tip

[114].

By using the equation above, the plot to log crack growth rate (da/dt) vs. strain energy release rate (G) is obtained.

In this study, subcritical crack growth was studied on epoxy/glass interfaces of untreated and y-APS treated glass which were aged in 85°C/85% RH. for 7 days. The results of untreated and y-APS treated samples are shown in Figure A3.1

and A3.2, respectively.

98

.........................

-

......

__

"

......

-

-

-

'

..

-

,-

.

-

.

.

. ..

-

.

-

-

1E-5 +-~-j--+--+---t-j-+-+--+-

..........

......

-

..........

.

-..

-...

".

......

_

--..

..........

.

".

'

'-_.'._

_

..

"._

_

_

...........................................................

_

_

..

.

-

_

._

.

-

-

'

..

'

..

.

.

.

-

'

..

' . '

'

..

..........

.

' . ' ' "

•....................

'

1E-6 +-~-j--+--+---t-t--+-+--+-

..........

..........

..........

'

_

'._

_

_

'

..

.

-

'

..

.

7.0 7.2 7.3 7.5 7.7 7.8 8.0 8.2 8.3 8.5 8.7 8.89.09.29.39.59.79.810.0

G

Figure A3.1: Plot of log crack growth rate (da/dt) in units of (mls) vs. strain energy release rate (G) in units of J/m

2 for untreated samples.

As seen from the plot, the crack growth rate did not really depend on G, which is easily seen from the very steep slope. Untreated sample showed the crack growth rate in the range of 9 x 10-

7

-

2

X

10-

5 mls.

99

1E-5

-+--:-::__ :: -::-::-::-::__ :=.::-0-: b:__ : :__ :1--::__ :----J:::

::: ::::: ::: ::::::::..:: ::: :::::::::::::::::: ::: :::: ::::::::::::: :::::::::::: :::::::

:::::::::::::::::1::: ::::::::::::::::::::: :::::::::::::::::::::::::::::::::::::::

1E-6 -I-:-:-':":":":"':":":":"---=::A-:-.__

-1-:-.__

•• __

.•__ ..,...,..,-1

~{.

1E-8

-:I-:-:':":":":"-:-flr..-:-:

..

.. __ .. __

:7.

.

••c:-: ••c:-: 1:-: c:-:

.

1E-9 - + - - - - - - i - - - - - - - + - - - - - - ; - - - - - - - - ;

20.0

21.3

22.5

23.8

25.0

G

Figure A3.2: Plot of log crack growth rate (da/dt) in units of (rn/s) vs. strain energy release rate (G) in units of J/m

2 for y-APS treated samples.

The similar result is shown in the case of y-APS treated samples. The plot shows a· steep slope but the crack growth rate is much slower than that of untreated sample. The crack growth rate was in the range of 2 x 10-

9

-

1

X

10-

5

mls. The G also shifts to the right due to the stronger adhesive strength at in interfaces. From the results, the subcritical energy release rate, G, is really close to the critical energy release rate, G c, so the critical

G testing is preferred since critical, G c tests are faster and easier to perform.

100

VITA

Orasa Khayankarn was born in Thailand on August 28, 1976. She graduated from

Hatyai Wittayalia High School in 1994 and entered the Prince of Songkhla University,

Thailand, the same year. She received the bachelor degree of Science in Chemistry with the second-class honor in 1998 before entering Chulalongkorn University for her master degree in polymer. She received the master degree in Polymer Science and Engineering from Chulalongkorn University in 2000 and was selected from the Royal Thai

Government to study at Lehigh University. In December 2000 she began her research at

Lehigh University under the supervision of Dr. Raymond A.

Pearson.

101

END OF

TITLE

Download