improved accelerated life testing for cathodic …

15
1 IMPROVED ACCELERATED LIFE TESTING FOR CATHODIC DELAMINATION by Kenneth M. Liechti and Thomas Mauchien Center for the Mechanics of Solids, Structures & Materials Aerospace Engineering & Engineering Mechanics University of Texas at Austin 210 East 24th Street, WRW 110C 1 University Station, C0600 Austin, TX 78712 512-471-4164 512-471-5500 (F) [email protected] CMSSM Report Number 2012-2 March 16, 2012

Upload: others

Post on 10-Apr-2022

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: IMPROVED ACCELERATED LIFE TESTING FOR CATHODIC …

1

IMPROVED ACCELERATED LIFE TESTING FOR CATHODIC DELAMINATION

by

Kenneth M. Liechti and Thomas Mauchien

Center for the Mechanics of Solids, Structures & MaterialsAerospace Engineering & Engineering Mechanics

University of Texas at Austin210 East 24th Street, WRW 110C

1 University Station, C0600Austin, TX 78712

512-471-4164512-471-5500 (F)

[email protected]

CMSSM Report Number 2012-2

March 16, 2012

Page 2: IMPROVED ACCELERATED LIFE TESTING FOR CATHODIC …

2

Introduction

The overall objective of this work is to establish the feasibility of modeling the cathodic delamination

problem in polymer coated submarine components with a view to developing a more accurate testing

standard for Accelerated Life Testing (ALT) and to determining the effectiveness of new approaches for

combating cathodic delamination in a quantitative manner. The importance of factors such as galvanic

potential, applied electrical potential, transducer materials, elastomeric materials, priming agents, non-

conductive coatings, etc. will eventually be considered as well as environmental conditions including

temperature, dissolved oxygen level, pH levels, salinity, etc.

Background

Cathodic delamination can be viewed as an example of environmentally assisted crack growth, an

example of slow, subcritical fracture that can eventually transition to fast fracture and the complete

failure of the interface. The environmental or chemical aspects of the problem have been studied for a

number of polymer/metal pairs and have the potential to be coupled with mechanical effects and

thereby improve our understanding of the problem. This in turn motivates accelerated life testing

protocols and the development of new surface treatments with a view to eliminating or at least

minimizing cathodic delamination in naval structures.

Cathodic delamination of polymer systems from metal substrates has been investigated frequently

because of its importance in automotive and naval systems. Several groups have used polybutadiene

films as model coating systems. Thus, Leidheiser et al. [1] reported on cathodic delamination of

polybutadiene films from steel substrates and found that high concentrations of OH- ions disrupted

polymer/substrate bonds. Koehler [2] investigated cathodic delamination of polybutadiene and

epoxy/phenolic films from steel; they also concluded that OH- ions disrupted polymer/substrate bonds

but also reported evidence for degradation of the films. Watts et al. [3] reported that cathodic

delamination of polybutadiene from steel originated in the oxide but eventually involved degradation of

the coating as well. Dickie et al. [4] investigated cathodic delamination of polybutadiene from steel; they

attributed failure of the coatings to hydrolysis of ester groups that were present in the cross-linked

polybutadiene. Watts and co-workers [5] also considered cathodic delamination of an epoxy/amine

coating from steel. They concluded that delamination initially involved reduction of the oxide but also

involved degradation of the coating. Dickie and co-workers [6] investigated cathodic delamination of

more complex systems, including an epoxy/urethane cross-linked with melamine/formaldehyde and an

epoxy/amine cross-linked with melamine formaldehyde or urea/formaldehyde. In all cases, they

attributed failure to degradation of the coating systems in the high pH environment created at the steel

surface during cathodic delamination.

Several groups have investigated cathodic delamination of neoprene rubber from steel substrates.

Stevenson [7] found that neoprene/steel bonds formed using proprietary primer and adhesive were

stable in an inert environment but degraded rapidly when cathodic potentials were applied to the steel.

Boerio and co-workers [8] investigated cathodic delamination of the same rubber/steel system and

suggested that inorganic chlorides formed in the bondline during curing of the rubber could lead to large

osmotic pressures in the bondline during exposure to water at high pH values and to failure of the

bonds. Boerio and Hong subsequently identified zinc chlorides as being especially important [9]. Kozinski

[10] showed that adding large amounts of an aminosilane "coupling agent" to the primer referred to

Page 3: IMPROVED ACCELERATED LIFE TESTING FOR CATHODIC …

3

above greatly reduced the rate of cathodic delamination of neoprene/steel bonds; the decreased rate of

failure was attributed to formation of amine hydrochlorides rather than soluble zinc chlorides during

cathodic delamination.

These chemical aspects of cathodic delamination are an integral component of environmentally assisted

fracture mechanics concepts. Traditionally this was used to address the slow growth of macroscopic

cracks in metals, ceramics and glasses in aggressive environments such as hydrogen, sea water and

moisture. In this approach, a pre-cracked specimen is loaded in the aggressive environment of interest

(e.g. water for glass), the crack growth is measured and the crack growth rate is correlated to a fracture

mechanics parameter such as the stress intensity factor K or strain energy release rate G .

Either one is a measure of the crack driving “force”, usually a function of the loading and crack

geometry. They can be thought of as “crack meters” that quantitatively gage the severity of cracks in

structural components under load. The fracture toughness

cK or toughness cG of a particular material is the critical

value of K or G that causes the fast crack growth that is

usually associated with final failure of cracked components.

Consequently, once the toughness of a material is known, it

is possible to predict what crack length and or load level will

give rise to fast fracture in a cracked component if the stress

analysis of potential crack paths has been conducted to

determine the variation of stress intensity factor with load

level and crack length.

Environmentally assisted crack growth is a form of slow or

subcritical fracture that occurs at stress intensity factors that

are less than cK . With reference to Figure 1 for moisture

assisted fracture in glass [11, 12], fast cracks are ones that

grow faster than mm/s at stress intensity factor levels above

0.8 MPaяŵ͘ ��ƚ�ƐƚƌĞƐƐ�ŝŶƚĞŶƐŝƚLJ�ĨĂĐƚŽƌ�ůĞǀ ĞůƐ�ďĞůŽǁ �Ϭ͘ϱ�D WĂяŵ͕ �ƚŚĞ�ĐƌĂĐŬ�ǀ ĞůŽĐŝƚLJ�ĚĞƉĞŶĚƐ�ŽŶ�ƌĞůĂƟǀ Ğ�

humidity and stress intensity factor and is considered to be controlled by reaction rates between water

vapor and strained Si-O bonds in the class. For some humidity levels there is a so-called plateau region

0.5 0.7K MPa√m, where the crack velocity is independent of stress intensity factor and only a

function of humidity. In this region, the crack growth is controlled by transport phenomena such as bulk

or surface diffusion and can depend on the viscosity of the environmental species. At stress intensity

factors greater than 0.7 MPaяŵ ͕ �ƚŚĞ�ĐƌĂĐŬ�ƟƉ�ŽƵƚƌƵŶƐ�ƚŚĞ�ƌĞĂĐƟŶŐ�ĞŶǀ ŝƌŽŶŵĞŶƚ�ĂŶĚ�ƚŚĞ�ƌĞƐƵůƚƐ�ĨƌŽŵ�Ăůů�

the relative humidity levels collapse onto one curve which is highly sensitive to the stress intensity

factor. Although the example given here is for glass in humid environments, similar features (chemical

reactions, diffusion controlled growth, etc.) have been observed in many metals [13], ceramics [14],

interfaces in microelectronics packages [15], adhesively bonded joints [16].

Such an approach has also been taken to characterize cathodically delaminating rubber/metal interfaces

[17, 18] and forms the basis for the current attempt to model cathodic delamination between

polyurethane and titanium. Once this model is established it will allow for an accelerated life protocol to

Fig. 1: Environmentally assisted

fracture of glass [11, 12].

Page 4: IMPROVED ACCELERATED LIFE TESTING FOR CATHODIC …

4

be developed where stress and temperature will be used to accelerate crack growth in laboratory

specimens while retaining the same crack growth mechanisms that are seen in service.

The elements of the approach are threefold:

1. Determining the mechanical behavior of the polyurethane and titanium. The former is more

challenging due to its nonlinearly elastic behavior.

2. Conducting a stress analysis of the specimen to be used in the cathodic delamination

experiment in order to design it to provide the anticipated range of energy release rate

values.

3. Conducting the cathodic delamination experiments and determining the crack velocity

profiles for the polyurethane/titanium interface as a function of energy release rate and

temperature.

A description of the work done in each of these areas is presented below.

Stress-Strain Behavior

The interface that was considered in this study was formed

by bonding CONATHANE® EN4/EN-7 polyurethane, a two-

component, non-MbOCA-based system, to titanium 6Al-4V

substrate using PR420 metal primer. The titanium was taken

to be linearly elastic with Young’s modulus and Poisson’s

ratio of 70 GPa and 0.32, respectively. The thickness of the

primer is so small that it does not need to be considered in

the stress analysis of the specimen, so the main emphasis

here is the uniaxial stress-strain behavior of the

polyurethane.

The tensile coupons were cast and cured to the geometry

shown in Figure 2 using manufacturer recommended

procedures. They were placed in an electromechanical

testing device and loaded to failure under displacement

control at a rate of 0.42 mm/s. The load and overall

extension of the specimen were measured. The engineering

stress in the specimen was obtained from the load and

original cross section area and the strain was the change in

length divided by the original length.

The results are shown in Figure 3, where the usual highly nonlinear response of an elastomer can be

seen. The results are compared with two linearly elastic approximations (50 and 5 MPa Young’s

modulus, both with Poisson’s ratio of 0.4) for the behavior of the polyurethane and the rubber that was

used in a previous study [18]. The red data points were used to represent the data (blue points) that was

actually taken during the experiment.

Three models of nonlinearly elastic behavior were considered: Mooney-Rivlin [19], Ogden [20] and

Marlow [21]. Each has its form of the strain energy function, which is then used to determine the stress

Fig.2: The geometry of the tensile

coupon that was used to determine

the stress-strain behavior of

EN4/EN7.polyurethane.

Page 5: IMPROVED ACCELERATED LIFE TESTING FOR CATHODIC …

5

strain behavior. For our purposes here, the polyurethane is assumed to be isotropic and incompressible.

The latter issue can be explored further in subsequent work.

0

2

4

6

8

10

12

0 1 2 3 4

Data

E=5MPa

E=50 MPa

Fitting data

[18]

Str

ess

(MP

a)

Strain

Figure 3: The stress-strain behavior of EN4/EN7 polyurethane.

The strain energy function of the Mooney-Rivlin material that was used here is

10 1 01 2( 3) ( 3)U C I C I , (0.1)

where 10C and 01C are material parameters, 2 2 21 1 2 3I and 2 2 2

2 1 2 3I are the first

and second deviatoric strain invariants, 1/3i iJ are the deviatoric stretches, i are the principal

stretches and J and eJ are the total and elastic volume ratios, respectively. The Ogden form of the

strain energy density function was

1 2 321

2i i i

Ni

i i

U

, (0.2)

where i , i , iD and N are material parameters. The Marlow form was

1e

d vU U I U J , (0.3)

where dU and vU are the deviatoric and volumetric components of the strain energy density function.

The best fits of these models to the measured stress-strain behavior are shown in Figure 4. For the

Mooney-Rivlin representation, 10C and 01C were 0.547 MPa and 2.272 GPa, respectively. The Ogden

parameters were:

Page 6: IMPROVED ACCELERATED LIFE TESTING FOR CATHODIC …

6

i (MPa) i

1 0.02 4.78

2 68.44 -2.02

The closest fit was provided by the Marlow model, which only deviated from the data at very small

strains. As a result, the Marlow was incorporated in subsequent stress analyses of the fracture

specimen.

Figure 4: A comparison of the Mooney-Rivlin, Ogden and Marlow models of the stress-strain behavior of

EN4/EN7 polyurethane with the data from the uniaxial test.

Fracture Specimen

The fracture specimen that

was used in this phase of the

work was the strip blister

specimen [18]. It had to be

redesigned for the

polyurethane and titanium

that were being considered

here. In view of the

nonlinearly elastic response of

the polyurethane, the analysis

was carried out with the finite

element code ABAQUS. The

specimen geometry is shown

in Figure 5. The specimen consisted of two polyurethane layers that were cast onto a titanium bar. A

Figure 5: The geometry of the strip blister specimen with a width

(not shown).

Page 7: IMPROVED ACCELERATED LIFE TESTING FOR CATHODIC …

7

central region was masked off during the curing to produce an initial crack of length 2a . A dowel of

diameter d was inserted along the centerline to provide a wedge loading to the specimen.

A plane strain formulation was used in

the finite element analysis and the

dowel was modeled as a rigid and

frictionless circular cylindrical surface.

The polyurethane and titanium were

given the stress-strain properties that

were discussed earlier. The deformed

shape of the specimen and the

distribution of the von Mises stress

with the dowel fully inserted are

shown in Figure 6. The blunted crack

with the highly deformed polyurethane

is clear to see. This was also the most

highly stressed region.

As indicated in the introduction,

fracture parameters such as the stress-

intensity factor and energy release rate are used to quantify the severity of cracks. Another parameter,

the J-integral is ideally suited as a fracture parameter for nonlinearly elastic materials, so it was

extracted from the finite element solutions. It was first used as a metric for determining the optimum

degree of mesh refinement or convergence of the solutions from the stress analyses. The results are

shown in Figure 7, where it can be

seen that mesh sizes less than 1 mm

provided converged solutions. The

finite element solution was checked

against previous results for the rubber

considered in [18] and was in close

agreement.

The strip blister specimen that was

considered in this work was 15.2 cm

long by 12.7 mm wide with

polyurethane layers that were 8 mm

thick. The initial crack length was 25.4

mm. A series of analyses were

conducted to determine the variation

of the J-integral with crack length and

dowel diameter. The results are shown

in Figure 8 where it can be seen that

the J-integral decreased sharply with

increasing crack length, making this a

stable crack growth configuration in the sense that the driving force for continued growth decreases as

Figure 6: The deformed shape of the strip blister specimen

and the distribution of the von Mises stress with the dowel

fully inserted.

Figure 7: A mesh refinement study based on the J-integral.

Page 8: IMPROVED ACCELERATED LIFE TESTING FOR CATHODIC …

8

the crack grows. Larger pin diameters increased the wedge effect and provided more energy for crack

growth.

Cathodic Delamination

This section describes the

specimen fabrication, the protocol

used in conducting the

experiments as well as the raw

data and data processing to

produce the crack speed vs. J-

integral plots, which are the basis

of the accelerated testing

procedure that is being proposed

here.

Specimen fabrication

Strip blister specimens were

fabricated according to the

geometry given in Figure 5 with

the same dimensions that were

used in the crack analysis. The

fabrication steps were as follows:

1. mask the 25 mm wide initial

crack area with mylar and

masking tape

2. grit blast the areas beyond the

no stick area with 36 grit alumina

3. quickly remove tape and place in MEK.

4. ultrasonic clean in two baths (1st removes major debris, 2nd final clean)

5. just as the part dries apply PR-420 metal primer to the bond surfaces

6. clean off any PR-420 from the initial crack area.

7. paint on PTFE (no silicone) release agent to the no stick area

8. apply mold walls

9. mix EN-7/EN-4, evacuate

10. fill mold to produce 7.94 mm layer, cure at 150°F for 16 hours

11. remove mold and repeat process on the opposite side

This process resulted in a specimen with two bondlines; one cured for 16 hours, the other for 32. Each

bondline had two crack fronts which were measured and averaged.

Dowel pins with a 4.8 mm diameter were inserted between the polyurethane and titanium at both

interfaces at the center of the initial crack and allowed to relax for several hours prior to environmental

exposure in order to minimize any viscoelastic effects.

Figure 8: J-integral solutions for all pin diameters. A Marlow

material model was used for the 8 mm thick polyurethane

layers.

Page 9: IMPROVED ACCELERATED LIFE TESTING FOR CATHODIC …

9

Protocol

The conditions for cathodic delamination were provided by placing the preloaded specimens in an

artificial sea water solution [22] in an aquarium and applying a suitable electrical potential to the

specimen. Air was directly injected into the artificial sea water using a Bubble Bar®, which, in

conjunction with a ventilator, ensured that there was a homogeneous and sufficient supply of oxygen

for the cathodic delamination reaction to occur. The artificial sea water was maintained at several

temperature levels (90.7, 86.0 and 77°F) in separate tests. A potentiostat was used to maintain the

metal parts of the specimen at a constant potential of -0.9 volts with respect to a standard calomel

electrode (SCE). A graphite rod was used as the anode. The current flow in the tank was measured

periodically throughout each experiment and remained constant at about 1mA. The specimens were

periodically removed in order to measure the crack length with vernier calipers. The crack length

measurements were made on both edges of the crack fronts and the average for each interface (16 and

32 hour cure) was recorded as the effective crack length.

Results

The crack growth histories for both interfaces and the three temperatures are shown in Figure 9. The

early growth appears in Figure 9a, starting from the initial crack produced by the masking. There was a

noticeable increase in crack length initially, which tapered off as the J-integral and crack driving force

decreased with increasing crack length. The greatest amount of crack growth occurred along the 16h

interface at the highest temperature. At each temperature, the 16h interface consistently gave rise to

more growth and the least amount of growth occurred at the lowest temperature. The amount of crack

growth decreased significantly beyond 500 hours (Fig. 9b), but continued until the tests were

discontinued after about 2 days.

Figure 9a: Early crack growth history along both interfaces (16h and 32h) at 90.7, 86.0 and 77°F.

30

35

40

45

50

55

0 20 40 60 80 100

86.0 86.0

77.0 77.0

90.7 90.7

2a

(mm

)

t (min)

32h (oF) 16h

Page 10: IMPROVED ACCELERATED LIFE TESTING FOR CATHODIC …

10

Figure 9b: Complete crack growth history along both interfaces (16h and 32h) at 90.7, 86.0 and 77°F.

Ultimately, the resistance to cathodic delamination is presented as crack growth rate as a function of

crack driving force, represented by the J-integral. In preparation for taking the derivative of the data in

Figure 9, several fitting schemes were considered. In the end, a least squares fit of the data power law of

the form 0na a t worked very well in all cases as demonstrated (Fig. 10) by examples of fits to the

data at 77°F in both the short and long term.

30

35

40

45

50

55

0 500 1000 1500 2000 2500

86.0 86.0

77.0 77.0

90.7 90.7

2a

(mm

)

t (min)

32h (0F) 16h

Page 11: IMPROVED ACCELERATED LIFE TESTING FOR CATHODIC …

11

Figure 10a: Power law fits to early crack growth data along both interfaces (16h and 32h) at 77°F.

Figure 10b: Power law fits to the entire crack growth history along both interfaces (16h and 32h) at

77°F.

30

35

40

45

50

55

0 20 40 60 80 100

32h 32h

16h 16h

2a

(mm

)

t (min)

Data Fit

30

35

40

45

50

55

0 500 1000 1500 2000 2500

32h 32h

16h 16h

2a

(mm

)

t (min)

Data Fit

Page 12: IMPROVED ACCELERATED LIFE TESTING FOR CATHODIC …

12

The resistance to cathodic delamination is shown in Figures 11a and b, which respectively display the

results of differentiating the power law fit to the delamination data and via a simple difference scheme.

At all temperatures, the 16h interfaces were less resistant to cathodic delamination than their 32h

counterparts although this was not quite as marked at 86°F. Each condition exhibited a threshold in J-

integral below which there was no growth. Threshold levels were dependent on temperature. Infinite

lifetime designs would attempt to ensure that the J-integral levels associated with any preexisting

delaminations in the service application fall below the threshold level. Above the threshold, the J-

integral had a strong effect on delamination growth rates. Consequently, it can be seen that both

temperature and stress level can be used to accelerate delamination in this regime. The discretely

differentiated data showed some signs of reaching a so-called plateau level above 2 kJ/m2, where

delaminations rates were independent of J-integral level. There were indications that the delamination

rates in the plateau region were independent of temperature.

Accelerated Life Testing

The preceding sections have demonstrated the basis for improved accelerated life testing of

polymer/metal interfaces that are subjected to cathodic delamination. The next step is to consider some

typical components to determine what range of J-integral values would be associated with

delaminations that are observed in service. If these were to fall in the range 0.8 2J J/m2, for

applications using this primer, then they would be candidates for this approach. Other surface

treatments would likely have different ranges of active driving forces, so an added benefit of this

approach is that the effectiveness of different surface treatments can be quantified, not only in terms of

data such as that presented in Figure 11 but also in predictions of anticipated lifetimes of interfaces with

preexisting flaws. This latter claim is based on being to conduct stress analyses of “hot spots’” in a

particular component and determine the J-integral values associated with potential delamination paths.

The resistance data (e.g. Fig.11) can then be integrated in order to predict how long it would take each

delamination path to reach critical values of J-integral. Predicted failure times could then be obtained as

a function of surface treatment, which would in turn allow cost optimizations to be made in a

quantitative manner.

Conclusions

A fracture mechanics approach to accelerated life testing of cathodic delamination between titanium

and polyurethane has been demonstrated. The approach recognizes that both temperature and stress

can be used to accelerate the process. It provides a quantitative and rational basis for conducting

accelerated testing as well as evaluating new surface treatments.

Page 13: IMPROVED ACCELERATED LIFE TESTING FOR CATHODIC …

13

(a)

(b)

Figure 11: Resistance to cathodic delamination of a titanium/polyurethane interface in ASW.

10-5

10-4

10-3

10-2

10-1

100

101

102

0.8 0.9 1 2 3

86.0

86.0

77.0

77.0

90.7

90.7

da

/dt

(mm

/min

)

J (kJ/m2)

32h (oF) 16h

10-5

10-4

10-3

10-2

10-1

100

101

102

0.8 0.9 1 2 3

77.0 77.0

86.086.0

90.7 90.7da

/dt

(mm

/min

)

J (kJ/m2)

32h (oF) 16h

Page 14: IMPROVED ACCELERATED LIFE TESTING FOR CATHODIC …

14

References

1. Leidheiser, H. and W. Wang, Some substrate and environmental influences on the cathodicdelamination of organic coatings. Journal of Coatings Technology, 1981. 53: p. 77-84.

2. Koehler, E.L., The mechanism of cathodic disbondment of protective organic coatings - aqueousdisplacement at elevated pH. Corrosion, 1984. 40: p. 5-8

3. Watts, J.F. and J.E. Castle, The Application of X-ray Photoelectron Spectroscopy to the Study ofPolymer-to-Metal Adhesion: Part 1: Polybutadiene Coated Mild Steel. Journal of MaterialsScience 1983. 18: p. 2987-3003.

4. Dickie, R.A., J.S. Hammond, and J.W. Holubka, Interfacial Chemistry of the Corrosion ofPolybutadiene-Coated Steel. Ind. Eng. Chem. Prod. Res, 1981. 20: p. 339-343.

5. Watts, J.F. and J.E. Castle, The Application of X-ray Photoelectron Spectroscopy to the Study ofPolymer-to Metal Adhesion: Part 2: The Cathodic Disbandment of Epoxy Coated Mild Steel.Journal of Materials Science 1984. 19: p. 2259-2272.

6. Holubka, J.W., et al., Surface Analysis of Interfacial Chemistry in Corrosion-Induced Paint LossAdhesion. Journal of Coatings Technology, 1980. 52: p. 63-68.

7. Stevenson, A., The effect of electrochemical potentials on the durability of rubber metal bonds insea-water. Journal of Adhesion, 1987. 21: p. 313-327.

8. Boerio, F.J., et al., Cathodic Debonding of Neoprene from Steel. The Journal of Adhesion, 1987.23(2): p. 99 - 114.

9. Boerio, F.J. and S.G. Hong, DEGRADATION OF RUBBER-TO-METAL BONDS DURING SIMULATEDCATHODIC DELAMINATION. Journal of Adhesion, 1989. 30(1-4): p. 119-134.

10. Kozinski, S.E. and F.J. Boerio, EFFECT OF AN AMINOSILANE ON THE CATHODIC DELAMINATIONOF ADHESIVE BONDS BETWEEN NEOPRENE AND STEEL. Journal of Adhesion Science andTechnology, 1990. 4(2): p. 131-143.

11. Wiederhorn, S.M., Influence of water vapor on crack propagation in soda-lime glass. AmericanCeramic Society -- Journal, 1967. 50(8): p. 407-414.

12. Freiman, S.W., S.M. Wiederhorn, and J.J.J. Mecholsky, Environmentally Enhanced Fracture ofGlass: A Historical Perspective. Journal of the American Ceramic Society, 2009. 92(7): p. 1371-1382.

13. Lynch, S.P., Failures of structures and components by environmentally assisted cracking.Engineering Failure Analysis, 1994. 1(2): p. 77-90.

14. Wiederhorn, S.M., E.R. Fuller Jr, and R. Thomson, MICROMECHANISMS OF CRACK GROWTH INCERAMICS AND GLASSES IN CORROSIVE ENVIRONMENTS. Metal science, 1980. 14(8-9): p. 450-458.

15. Dauskardt, R.H., Lane, M., Ma Q., and Krishna, N., Adhesion and debonding of multiplayer thinfilm structures. Eng. Fract. Mech., 1998. 61: p. 141-162.

16. Ameli, A., M. Papini, and J.K. Spelt, Hygrothermal degradation of two rubber-toughened epoxyadhesives: Application of open-faced fracture tests. International Journal of Adhesion andAdhesives. 31(1): p. 9-19.

17. Liechti, K.M., and Adamjee, W., Mixed-mode Cathodic Delamination of Rubber from Steel. 1992.40: p. 27-45.

18. Liechti, K.M., et al., A FRACTURE-ANALYSIS OF CATHODIC DELAMINATION IN RUBBER TO METALBONDS. International Journal of Fracture, 1989. 39(1-3): p. 217-234.

19. Mooney, M., A Theory of Large Elastic Deformation. J. Appl. Phys., 1940. 11(582-592).

Page 15: IMPROVED ACCELERATED LIFE TESTING FOR CATHODIC …

15

20. Ogden, R.W. Large Deformation Isotropic Elasticity - On the Correlation of Theory andExperiment for Incompressible Rubberlike Solids. in Proc. Royal Soc. London. 1972.

21. Marlow, R.S. A general first-invariant hyperelastic constitutive model. in Constitutive Models forRubber III: proceedings of the Third European Conference on Constitutive Models for Rubber.2003. London: Taylor & Francis.

22. Petco. http://www.petco.com/product/101391/Petco-Premium-Marine-Salt-Mix.aspx 2012.