Você está na página 1de 28

4 Jun 2003 12:24 AR AR189-MR33-02.tex AR189-MR33-02.

sgm LaTeX2e(2002/01/18) P1: IKH


10.1146/annurev.matsci.33.012202.130440

Annu. Rev. Mater. Res. 2003. 33:29–54


doi: 10.1146/annurev.matsci.33.012202.130440
Copyright ° c 2003 by Annual Reviews. All rights reserved
First published online as a Review in Advance on January 24, 2003

INTERFACE FRACTURE
Michael Lane
IBM T. J. Watson Research Center, Yorktown Heights, New York 10598;
email: mwlane@us.ibm.com

Key Words adhesion, macroscopic work of fracture, delamination, debonding,


work of adhesion
■ Abstract Interfacial adhesion plays a central role in a number of technologically
important applications. Quantitatively measuring the adhesion of an interface and un-
derstanding the processes and controlling mechanisms of energy dissipation is not
always a straightforward task, however. It is often not enough to know that an interface
in a particular application is weak and prone to failure because it can be difficult to
accurately recreate that interface with a bulk specimen and derive a meaningful adhe-
sion value. Rather, a fundamental knowledge of the processes that actually contribute
to the interfacial strength is important, so that, when bulk specimens are prepared,
care can be taken to eliminate energy-absorbing processes that are not present in the
actual application. Accordingly, this paper reviews some of the literature highlighting
the contributing factors that control interfacial adhesion. The focus is on those models
that describe the detailed mechanisms of the energy-absorbing processes and on some
of the experimental data that illustrates those models.

INTRODUCTION

Interface Adhesion Versus Macroscopic Work of Fracture


The energy associated with the separation of an interface is often referred to as
the macroscopic work of fracture, 0 i. Contributions to this overall energy arise
from a number of different processes that can generally be placed in one of two
groups (Figure 1). The first group is the work of adhesion, 0 o, which is the energy
required to separate the interface of interest or, more basically, to rupture bonds at
the interface. The work of adhesion leverages all other energy-absorbing processes
(1, 2). Because of this, small changes in interfacial bond strength can result in much
larger changes in the macroscopic work of fracture. The second group is that of the
energy-absorbing processes that scale with the work of adhesion. These processes
include irreversible contributions due to surface creation (3, 4), and plastic energy
dissipation (2, 5–7), and crack face interactions such as asperity contact and liga-
ment bridging (8). Each of these terms has been described as having a multiplicative
effect on the work of adhesion and may be summarized as follows (1–4, 8):

0084-6600/03/0801-0029$14.00 29
4 Jun 2003 12:24 AR AR189-MR33-02.tex AR189-MR33-02.sgm LaTeX2e(2002/01/18) P1: IKH

30 LANE

Figure 1 Schematic representing energy dissipation processes that contribute


to the macroscopic work of fracture (17).

0i = (ξ + χ + ζ )0o , 1.

where ξ is the irreversible contribution of surface creation having an approximate


value between 1 and 10 (3, 4); χ is a roughness-related shielding parameter that
scales as the amplitude of the roughness squared and inversely as the wavelength of
the roughness (8); and ζ characterizes the plastic energy dissipation and scales with
the square of the yield stress (2). Although Equation 1 is instructive, the detailed
models that have been developed to describe these processes are key to under-
standing the mechanisms that control the macroscopic work of fracture. However,
before discussing these mechanisms in detail, we briefly review the effect of stress
state on interfacial adhesion and cite some commonly used quantitative adhesion
techniques, focusing on their origin, strengths, and weaknesses.
4 Jun 2003 12:24 AR AR189-MR33-02.tex AR189-MR33-02.sgm LaTeX2e(2002/01/18) P1: IKH

INTERFACE FRACTURE 31

Figure 2 Schematic representing the increase in the


macroscopic work of fracture expected as the phase angle
of loading is increased (2).

Effect of Stress State on Interfacial Adhesion


The macroscopic work of fracture is non-unique because it depends on the phase
angle of loading, which is related to the ratio of shear to normal stresses at the
crack tip (9–13). Thus any adhesion measurement must be reported relative to
its phase angle of loading. A schematic illustrating the relationship between 0 i
and the applied phase angle is shown in Figure 2 (2). The increase in 0 i may
be attributed to greater contributions of plasticity and debond face interactions
such as roughness-induced shielding (2, 13). It is important to note that due to the
elastic modulus mismatch at a bimaterial interface, stress fields at the debond tip
are rotated from those expected from the far-field applied loading so that some
degree of shear stress is always acting on the debond tip. The relative degree of
mismatch is determined by the Dundurs’ parameters (14), and detailed papers on
the mechanics, as related to stress field rotation (2, 13) and crack path selection
(15), have dealt with these topics in much greater detail than could be accomplished
here. See the excellent review article on this subject by Hutchinson & Suo (13).

Adhesion Measurement Methods


Although a number of tests have been devised to measure interfacial adhesion,
only those that give quantitative results are useful in elucidating controlling mech-
anisms at interfaces. Three of the most accessible techniques which have been
used extensively in recent literature (e.g., 16–18) are briefly discussed here with a
focus on identifying the strengths and weaknesses of each. These techniques are
the indentation, superlayer delamination, and sandwiched fracture mechanics tests
with schematics of each shown in Figure 3.
4 Jun 2003 12:24 AR AR189-MR33-02.tex AR189-MR33-02.sgm LaTeX2e(2002/01/18) P1: IKH

32 LANE

Figure 3 Common quantitative techniques for measuring interfacial adhesion:


(a) indentation test, (b) super layer delamination, and (c) sandwiched fracture
mechanics specimens.

Indentation tests (Figure 3a) are characterized by the use of a sharp-rigid tip
indenting a film (or films) on a rigid substrate. The original mechanics for a single
film were derived by Marshall & Evans (19) and indicate that the driving force for
delamination, or the strain energy release rate, G, has three contributing energies
that are required for its calculation. The required measurements are the residual
stress of the film, the indentation depth that allows for calculation of the indentation
stress, and the diameter of the resultant delamination. One advantage of this test is
that the interface of interest undergoes no processing after its initial creation, which
allows for the determination of the interfacial adhesion of the virgin interface.
However, for ductile films or well-adhering interfaces, the film of interest may
deform plastically prior to the buildup of sufficient elastic strain energy, resulting
in no delamination (20). In order to address this issue, a highly stressed overlayer
film may also be deposited (21). The overlayer adds an additional driving force
through its residual stress and is typically a refractory metal with a high-yield stress
that reduces plastic deformation. Kriese & Gerberich (21) extended the analysis
of indentation of a single film to include the case of multiple films on a single
substrate. Interface adhesion measurements utilizing the superlayer indentation
test have captured effects of interfacial chemistry and plasticity (22–24), but the
scatter in the data collected can be large [at times up to an order of magnitude
(18)]. A further limitation of the indentation test is that methods to measure the
4 Jun 2003 12:24 AR AR189-MR33-02.tex AR189-MR33-02.sgm LaTeX2e(2002/01/18) P1: IKH

INTERFACE FRACTURE 33

effects of the environment, such as moisture-assisted debonding, have not been


developed (see below).
Utilization of a stressed superlayer (Figure 3b) to propagate an interface delam-
ination was first described by Bagchi et al. (25). The essential features of the test
include thin strips containing the interface of interest, a highly stressed overlayer,
and a release layer, serving as a precrack or weakly adhering region that eases
debond initiation. The driving force for delamination is the residual stress in both
the film of interest and the highly stressed superlayer film. The original analysis
of the delamination treated it as fully elastic and therefore did not account for any
plastic deformation in the films, which in certain instances may yield results that
do not accurately reflect the interface fracture energy. Also, the interface fracture
energy is found when the superlayer thickness exceeds a critical value, making
the preparation of a number of samples necessary and limiting the expediency
of the test. Xu et al. (26, 27) addressed these issues and found that the interface
fracture energy could be determined by a single measurement using microstrips
having a superlayer that exceeds the critical thickness by altering the geometry
from a single decohering strip to one that resulted in converging decohesion strips
(16, 26). The analysis was extended beyond its elastic case to account for plasticity
through finite element calculations, and the technique can be used to evaluate envi-
ronmental effects on delamination (16, 26, 27). The need to calibrate each sample
structure with finite element method (FEM) calculations does place some limits on
the accessibility of the technique, however. Further complications can arise if the
decohered strips interfere with themselves and result in frictional losses. However,
this should not be considered a major limitation of the test, as this occurrence
is in general easily observable and the data from those samples can be excluded
(16).
A number of sandwiched fracture mechanics specimens exist (28–31), but a
general feature is that the interface of interest is sandwiched between two large
elastic substrates (Figure 3c). By doing this, the residual stresses in the debonding
films are constrained and therefore do not contribute to the energy that drives the
delamination. In certain configurations, a small contribution to the overall driving
energy for delamination, from the elastic curvature of one substrate, occurs but is
typically three orders of magnitude less than the measured adhesion values (17).
Of the various sample geometries that can be made, two are discussed here: the
four-point flexure and the double-cantilever beam (DCB). The four-point flexure
geometry first described by Charalambides et al. (31) is the simpler of the two
specimens in that the strain energy release rate is a function of only the specimen
geometry and the critical load for debond growth. Because the debond length does
not enter into the calculation, it does not have to be measured, greatly simplifying
the testing. However, the strain energy release rate for the DCB geometry (29)
depends on both the load and the debond length, making the data collection more
problematic because the debond length can be difficult to determine accurately in
small samples. The phase angle applied to the interface in the two specimens differs
from ∼43◦ , for the four-point flexure, to nominally 0◦ , for the DCB geometry.
4 Jun 2003 12:24 AR AR189-MR33-02.tex AR189-MR33-02.sgm LaTeX2e(2002/01/18) P1: IKH

34 LANE

However, the DCB sample can be placed in a special loading configuration (32–
34) that allows the interface fracture energy to be determined over the complete
range of phase angles, as has been done in a number of glass epoxy systems
(35–38). Load relaxation techniques can be used to measure the effects of the
environment (39), and the samples can also be used to address issues of fatigue
loading on the interface (40), making this particular test the most flexible reviewed.
However, there are several drawbacks to this test. First, the sandwiched fracture
mechanics specimens are often made either by a diffusion bonding procedure or
with an epoxy. The diffusion bonding procedure often involves high temperatures
for extended periods of times [400◦ C in the case of Cu diffusion bonding (17)]. This
introduces the possibility of altering the interface during the sample preparation
step. In the case of an epoxy bond, care must be taken to design the stack structure
so that plasticity in the bond layer is eliminated (41), otherwise measurements of
the interface fracture energy will include plasticity in the epoxy bond layer. The
sample preparation of the sandwiched fracture mechanics specimens can be tedious
and often requires polished side face of the samples if the substrate is brittle, e.g.,
Si, to eliminate fracture of the substrate before interfacial delamination occurs.
Finally, Becker et al. (42) have shown that, for the four-point flexure geometry, the
limit of validity of the mechanics-derived strain energy release rate depends on
the amount of plastic energy dissipation, which is a function of the ductile layer
thickness.

EFFECTS OF INTERFACIAL CHEMISTRY

The driving force for formation, and therefore the energy that must be overcome
during fracture, of an interface is the release of energy when intimate contact is
established between the materials that form the interface (43, 44). The simplest
description of the physical interaction between two materials is the work of adhe-
sion, 0 o. If clean surfaces are brought into contact, the energy is released according
to (e.g., 44):
0o = γ1 + γ2 − γ1–2 , 2.
where γ 1 and γ 2 are the free energies of the relaxed surfaces of each constituent
of the bimaterial interface, and γ 1–2 represents the energy of the relaxed interface.
Note this is simply the Griffth criteria for fracture (45) applied to interfaces, and 0 o
is the reversible work released per unit area of interface formed by two free surfaces.
This section discusses the factors that affect the work of adhesion and, in particular,
deals with the effects of interfacial segregants, bond density, and bond energy.
Rice & Wang (46) formally described the effect of segregants to an interface
on the work of adhesion as
X¡ ¢
Go = 0o − 1goi − 1gso ci , 3.
i

where Go is the work of adhesion modified by segregation, ci is the concentration of


4 Jun 2003 12:24 AR AR189-MR33-02.tex AR189-MR33-02.sgm LaTeX2e(2002/01/18) P1: IKH

INTERFACE FRACTURE 35

segregant per unit area of interface, and 1go is the Gibbs free energy of segregation
for the interface (i) and a free surface (s).
An example of segregation is seen in the effect of nitrogen on the adhesion of
the Ta/SiO2 interface (47, 48). In this experiment a 2-nm Ta layer was deposited
on thermal SiO2 and different amounts of N (20, 30, or 40 at%) were diffused
to the interface. This resulted in an increase in the measured fracture energy by
∼150% owing to the addition of the nitrogen, and the increase may be understood
by consideration of the following reactions:
14 Ta + 5 SiO2 → 2 Ta2 O5 + 5 Ta2 Si—43.9 kcal/mol, 4a.

26 Ta + 20 SiO2 + 10 N2 → 8 Ta2 O5 + 5 Ta2 Si + 5 Si3 N4 —609.6 kcal/mol, 4b.


where the change in free energy of the reaction is given to the right of the reaction.
As N moves to the interface, the reaction in Equation 4b is favored over that in 4a
because it is more energetically favorable. Because N is the limiting reagent, more
nitrogen leads to a more complete reaction at the interface. X-ray photoelectronic
spectroscopy (XPS) of the fracture surfaces and depth profiling of virgin interfaces
confirmed the increase in reaction products described by Equation 4b and found
that the absolute amount of reaction products increased as the N content at the
interface increased (Figure 4).
Of course, segregants to an interface can have deleterious effects on interface
fracture resistance, as evidenced by the work of Lipkin et al. (49). The authors found

Figure 4 Increase in interface debond energy measured as a function of interfacial


N. The inset shows XPS spectra corresponding to reacted Ta at the interface. As the N
content at the interface increases, the amount of reacted Ta increases (47, 48).
4 Jun 2003 12:24 AR AR189-MR33-02.tex AR189-MR33-02.sgm LaTeX2e(2002/01/18) P1: IKH

36 LANE

that ultraclean Al2O3/Au interfaces, formed by first evaporating Au onto Al2O3


and then diffusion bonding two identical samples together to create a standard
fracture mechanics specimen, and tested in dry environments, had interface fracture
energies that exceeded 250 J/m2. When the diffusion bonding was carried out in
the presence of a graphite powder, carbon was found to diffuse along the interface,
and the interface fracture energy was reduced to as low as 2 J/m2. The dramatic
decrease in interface fracture energy may be understood by noting that the carbon
segregation serves two purposes. The first is to reduce the work of adhesion, and
the second effect is a corresponding decrease in plasticity in the Au layer due to
the reduced work of adhesion. The relationship between the work of adhesion and
plasticity is discussed in more detail below, but the work of Lipkin et al. serves
as an excellent example of the dependence of plastic dissipation on the work of
adhesion.
More recently, Korn et al. (50) and Cannon et al. (51) have examined the
Nb/Al2O3 system and found that by altering the segregant to the interface, the in-
terface fracture energy can be either increased or decreased. Specifically, a reactive
segregant such as Ti was found to increase the interface fracture energy, whereas
nonreactive segregants such as S and Ag were found to have a deleterious effect.
This is consistent with both the N effect on the Ta/SiO2 interface and the C effect
on the Au/Al2O3 interface, as N is known to improve the toughness of bulk glasses
(and so it may have a similar function at the glass/Ta interface) and Au has no
affinity for C.
Segregation to, or chemistry of, an interface is not the only way to effect the
work of adhesion. Rather, changes in the number of interfacial bonds can dra-
matically affect the interface fracture resistance. In investigations of the work of
adhesion between an epoxy and a self-assembled monolayer (SAM), Zhuk and
coworkers (52) found that as the ratio of terminal groups or bonding sites on the
SAM was altered, a corresponding change in the work of adhesion occurred. By
varying the ratio of CH3 (nonbonding) to COOH (bonding) groups, the work of
adhesion was increased by >30%. Lin (53) found similar results for the effect of a
benzophenone adhesion promoter on the interface fracture energy of the PMMA/Si
system. Benzophenone is a photochemically activated adhesion promoter, in which
the functional groups are activated by exposure to ultraviolet light. By increasing
the UV exposure time, the number of bonding sites was found to increase and a
corresponding increase in the interface fracture energy was observed.
These results underscore the dependences of the work of adhesion on both the
chemistry of the interface and the number of bonds across the interface. Building
upon these ideas, Lane considered a simple reaction of the type:
A + B → C, 5.
where the change in free energy for the reaction is given by
1Grxn = 1GC − 1GA − 1GB , 6.
where 1G is the Gibbs free energy of component A and B and the interface C. If
4 Jun 2003 12:24 AR AR189-MR33-02.tex AR189-MR33-02.sgm LaTeX2e(2002/01/18) P1: IKH

INTERFACE FRACTURE 37

the nonequilibrium thermodynamic state of the surface created by fracture does


not contribute significantly to the energy dissipation process, ξ ∼ 1 in Equation 1,
and plasticity and crack face interactions are minimal, then the change in Gibbs
free energy is the work of adhesion (54). Considering thin metals deposited at
room temperature in UHV conditions on thermally grown SiO2, where interface
roughness effects are minimal and plastic dissipation is constrained, Lane showed
that the Gibbs free energy for a given family of metal oxides could be normalized
by the number of oxygen atoms to give a single value representative of the metal-
oxygen bond energy, and that a plot of this versus the interface fracture energy
resulted in a straight line with a slope approximately equal to the expected bond
density of oxygen atoms on a SiO2 surface (Figure 5). This indicates that for this
particular case, the work of adhesion may be represented by

0o = N (1g M + 1g O − 1g M – O ), 7.

where N is the number of bonds per unit area, and 1g is the Gibbs free energy
of the pure metal, the oxygen-terminated Si surface, and the metal-oxygen bond,
respectively. Equation 7 is similar to that derived for the fracture resistance of bulk
glasses in a corrosive environment (55) if one notes that the chemical potential
(used in the bulk glass literature) is equivalent to the partial molar Gibbs free
energy.

Figure 5 Relationship determined between the work of


adhesion of metal/SiO2 interfaces and the enthalpies of
formation determined from bulk oxide compounds (54).
4 Jun 2003 12:24 AR AR189-MR33-02.tex AR189-MR33-02.sgm LaTeX2e(2002/01/18) P1: IKH

38 LANE

EXTRINSIC ENERGY DISSIPATION PROCESSES


Plasticity
As has been observed in the case of the Au/Al2O3 interface, plastic energy dis-
sipation can be a large contribution to the interface fracture energy but one that
depends strongly on the work of adhesion of the interface. As a result, there is a
need to define the important parameters that control the work of adhesion at an
interface and relate those properties to plastic energy dissipation. Accordingly, de-
tailed models have been developed to describe the fracture process and relate them
to plastic energy dissipation (e.g., 7, 56–68). The purpose of this section is not to
give detailed reviews of these models but rather to discuss one or two that have
clearly defined interfacial parameters and review their effectiveness in modeling
experimental data and the trends that arise as a result. The model reviewed in the
most detail is the unified model (65), which combines two previous models, the
SSV [after Suo, Shih, and Varias (64)] and the EPZ [embedded process zone (7,
67)]. A second model combines a multiscale continuum/FEM approach based on
cohesive surface elements or virtual internal bonds (59) and is discussed briefly
because it has been used to fit experimental results over a wide range of ductile
layer thicknesses and because many of the modeling results generated by it have
been related to predictions in the unified model (69).
Before beginning, consider the modeling link scale given by (7):
1 E0o
Ro = , 8.
3π (1 − υ 2 )σys2

where E, ν, and σ ys are the elastic modulus, Poisson’s ratio, and yield strength of the
plastically deforming material, respectively. Ro scales with the plastic zone size.
The essential features of the unified model may be found in Figure 6 and
consist of a traction separation law, which determines the interfacial strength, the

Figure 6 Schematic representing the key features of the unified model. The major
dependencies are indicated in the text except for N, which is the strain-hardening
exponent for the ductile layer (65).
4 Jun 2003 12:24 AR AR189-MR33-02.tex AR189-MR33-02.sgm LaTeX2e(2002/01/18) P1: IKH

INTERFACE FRACTURE 39

peak stress σ̂ , and a plasticity free zone of width D. The traction separation law
characterizes the interface fracture process and serves as an internal boundary
condition along the interface. A dimensionless separation parameter, λ, is defined
in terms of the normal, δ n, and tangential, δ t, displacements of the crack faces
across the interface:
sµ ¶ µ ¶2
δn 2 δt
λ= + c , 9.
δnc δt
where δnc and δtc are critical interface displacements. When λ = 1, the tractions drop
to zero, crack advance occurs, and the fracture energy is determined by
1 c
0o = σ̂ δ [1 − λ1 + λ2 ], 10.
2 n
where λ1, λ2, σ̂ , and δnc are detailed parameters in the traction separation law and
may be found in References (62, 65). The delamination is separated from the active
plastic zone by an elastic layer of thickness, D, the purpose of which is to embed the
debond tip in an elastic region so that the high stresses at the debond tip necessary
for atomic separation can be obtained (64). Therefore the interface is characterized
by the adhesion energy and peak separation stress, 0 o and σ̂ , and the plasticity
free zone of width D. Simulation results indicate that the ductile layer deformation
properties and film thickness, cohesive properties, and elastic layer thickness have
the most significant effect on interface fracture resistance (e.g., 63, 65).
A problem in determining the effect of any one parameter on the extent of plastic
energy dissipation in a given system is the number of independent variables in the
model. In order to accurately gauge the effect of any one variable on a given system,
the others must be fixed by experimental measurements. One attempt to do this is
found in the experimental study of the fracture resistance of the TiN/SiO2 interface
with different thicknesses of a ductile Al-Cu layer (17, 48, 69). In this experiment,
the thickness of the elastic layer, D, was set by the thickness of the TiN layer,
and an estimate of the work of adhesion was made by extrapolation to zero Al-Cu
layer thickness. This left only the peak interface separation stress and the yield
properties of the Al-Cu layer as variables. The results of the interface fracture
energy measurements along with simulations of the fracture process from both the
unified model (65) and a multiscale continuum/FEM model (63) are presented in
Figure 7. Both models accurately predict the observed behavior, and an important
ratio emerges from the unified model. As the ratio of the ductile layer thickness,
hAl-Cu, to the modeling length scale (given in Equation 8, Ro ∼ 1 µm in the current
example) approaches four, increases in toughening owing to additional increases in
metal layer thickness are not expected. Thus for a fixed value of D and metal yield
stress, limits to the toughness enhancement owing to plastic dissipation can be
expected, as was predicted (65). However, the utility of the simulations is limited
because accurate values of the work of adhesion, 0 o, were not determined and a
constant value of the yield stress of the Al-Cu layer was assumed, even though the
thickness of the film varied by over an order of magnitude. Note that data from
4 Jun 2003 12:24 AR AR189-MR33-02.tex AR189-MR33-02.sgm LaTeX2e(2002/01/18) P1: IKH

40 LANE

Figure 7 Interface debond energy as a function of Al-Cu layer thickness. The lines
represent fits of various models to the data. In the rising portion of the curve, the plastic
zone (represented by a black rectangle) size is larger than the Al-Cu layer thickness.
The plateau region indicates the plastic zone size has become smaller than the Al-Cu
layer thickness (17, 48, 69).

the double-cantilever beam geometry are also presented and illustrate the effect of
stress state on the plastic deformation in these samples (depicted schematically in
Figure 2) (48).
However, measurements of interface fracture energy have been made for ductile
metal films where the interface fracture energy did not change over a range of metal
layer thickness (18, 48, 69), which indicates that the constant interface fracture
energy is a measure of the work of adhesion of the interface, as there is no apparent
plastic energy dissipation. One of these systems has been modeled extensively with
the multiscale continuum/FEM approach developed by Lane and Lane et al. (48,
69). The system consists of the SiO2/TaN interface with Cu layers ranging from
0.03 to 18 µm in thickness. Delamination occurred along the SiO2/TaN interface
so that the thickness of the elastic layer, D, is set by the thickness of the TaN layer,
and the work of adhesion is set by the plateau value of the interface fracture energy
shown in Figure 8. This leaves only the peak interface separation stress, σ̂ , and
the metal yield stress as variables to fit the data. However, published Cu yield data
were used, leaving only the peak interface separation stress as a variable. For Cu
thicknesses where yield data were not available, the metal yield stress was used as
4 Jun 2003 12:24 AR AR189-MR33-02.tex AR189-MR33-02.sgm LaTeX2e(2002/01/18) P1: IKH

INTERFACE FRACTURE 41

Figure 8 Interface debond energy as a function of Cu layer thickness. The solid


line represents the fit to the data. The work of adhesion for the TaN/SiO2 interface is
determined by the plateau at thin Cu layer thicknesses. The plastic zone size exceeds
the Cu layer thickness even for the thickest films (48, 69).

a fitting parameter but made to follow a Hall-Petch type relation (69). The solid
line in Figure 8 shows the fit for a peak interface separation stress equal to ESi/60.
With σ̂ determined, other modeling parameters were varied to access their effects
on plastic energy dissipation. The results may be summarized as follows:

1. As σ̂ increases for a fixed value of 0 o, the interfacial toughening due to


plasticity increases consistent with predictions from the models (7, 63, 65).
2. For a fixed yield stress, σ ys, and ratio of Ro/D, there exists a critical ratio of
σ̂ /σ ys above which toughening due to plasticity is limited, as predicted in
(65, 69).
3. The interface fracture energy, 0 o, length scale ratio, Ro/D, and stress ratio,
σ̂ /σ ys, are the dominant factors in controlling the extent of plastic dissipation
(Figure 9), which has been predicted (62, 63, 65).

Effects of Interface Roughness


Energy dissipation due to interfacial roughness may manifest itself in a number of
ways, each of which will have its own dependence on the amplitude and wavelength
of the roughness. One possibility is that as the debond travels along the interface
4 Jun 2003 12:24 AR AR189-MR33-02.tex AR189-MR33-02.sgm LaTeX2e(2002/01/18) P1: IKH

42 LANE

Figure 9 Effects of interfacial parameters on the macroscopic work of fracture for


the TaN/SiO2 interface, as determined from the modeling results in Figure 8. Ro/10D
has been plotted for convenience rather than Ro/D (48, 69).

the debond tip propagation direction relative to the far-field applied loading may
change. This change in debond orientation results in a change of the near-tip stress
state and therefore a change in the debond driving force (70, 71). As the debond
moves it may either tilt or twist relative to its original path. The driving force for
the new orientation may be calculated by a tensor transformation of the original
far-field stress fields (72, 73). Plots of the tip driving force normalized by the far-
field driving force have been made for the special case of equal parts shear and
normal applied loading and serve as a useful guide for understanding the effect
(73). The results indicate that tilt or twist angles with a ratio of Gtip/Gapplied less
than 1 result in debond tip shielding, whereas ratios greater than 1 lead to debond
extension at driving forces reduced from that of a planar debond. Small changes
in propagation angle due to twist or tilt do not have a significant effect on the
macroscopic work of fracture.
The second increase in interface fracture energy due to nonplanarity may be
understood by consideration of the Griffith criterion for fracture, which indicates
the energy required to separate two surfaces depends on the total area of separation
(45). For a planar interface, this is simply the length of the debond times its width.
However, for a debond that deviates from a planar path, an increase in the total
debonded area occurs that results in a corresponding increase in interface fracture
energy that scales with the true debonded area over the planar area.
4 Jun 2003 12:24 AR AR189-MR33-02.tex AR189-MR33-02.sgm LaTeX2e(2002/01/18) P1: IKH

INTERFACE FRACTURE 43

The final mechanism to be considered is asperity contact behind the debond


tip, which acts to shield the debond from the applied loading (8). As the asperities
come into contact, normal tractions develop across the contact area, and frictional
forces must be overcome in order to pull the asperities apart. The effectiveness of
the shielding depends on the length of the contact zone, amplitude and wavelength
of the roughness, and the coefficient of sliding friction. Models that properly
account for each of these factors are very difficult to construct and implement in
a straightforward manner. Difficulties arise due to the various scales of roughness
that are inherent in nonplanar interfaces and the fact that the coefficient of sliding
friction is often not known. However, simplifications and assumptions about the
interface structure allow for the dependencies on the relevant interfacial parameters
to be determined. Evans & Hutchinson (8) proposed such a model in which an
idealized interface morphology, composed of one roughness and wavelength, is
examined.
A contact zone arises behind the debond tip and shields the tip from the applied
loads. Frictional effects are ignored so that the Mode I stress intensity at the debond
tip is equal to the applied Mode I stress intensity, which allows for determination
of the debond opening behind the debond tip (8). The effectiveness of the shielding
manifests itself in terms of a shielding parameter that relates the amplitude and
wavelength of roughness to the magnitude of the shielding, as given by

EH2
χ= , 11.
λ0o
where H is the amplitude of roughness, λ the wavelength, E the elastic modulus,
and 0 o the work of adhesion. A number of important phenomena are captured by
this simple model. First, the shielding effect increases dramatically as the loading
mode tends toward pure shear loading, as has been observed experimentally in a
number of systems. Second, when the shielding parameter is ≤10−4, shielding is
only realized as the phase angle of loading approaches 90◦ (pure shear). Finally,
as the shielding parameter approaches 1, further increases in the amplitude of
roughness do not significantly enhance the interface fracture energy.
Cao & Evans (10) found that for a glass epoxy system the interface fracture resis-
tance increased according to a value of χ equal to ∼1 as the phase angle of loading
was changed. Ma et al. (74) and Clarke (75) demonstrated that the interface fracture
energy of TiN/SiO2 could be increased by 50% simply by roughening the interface
(Figure 10). In a similar experiment examining the TiN/SiO2 interface, Lane et al.
(48, 76) examined a range of interfacial morphologies and correlated the change in
interface fracture energy with changes in interfacial roughness. In this experiment,
the interfaces were characterized with atomic force microscopy (AFM) to deter-
mine the average amplitude and wavelength of the roughness. A model interface
similar to that described by Evans & Hutchinson was constructed on the basis of
the root mean square (RMS) roughness and the average roughness wavelength, as
determined in the AFM measurements. These model interfaces were then used to
predict the relative magnitudes of the increases expected from the debond kinking
4 Jun 2003 12:24 AR AR189-MR33-02.tex AR189-MR33-02.sgm LaTeX2e(2002/01/18) P1: IKH

44 LANE

Figure 10 TiN/SiO2 interfaces with different degrees of interfacial roughness.


The rougher interface has 50% higher macroscopic work of fracture (74, 75).

out of the interface, increased surface area, and roughness-related shielding. Re-
sults indicated that increases in interface fracture energy for the roughest interface
predicted by crack tip deflection and increased area were only 2.3 and 0.3%, respec-
tively. However, increases based on the asperity contact model properly captured
the 25% increase in interface fracture energy as measured (Figure 11). So, espe-
cially for thin film systems where the roughness is not sufficient to cause large
changes in the propagation direction or significant increases in debonded area,
asperity contact is the dominant mechanism that leads to increases in interface
fracture energy due to roughness.

ENVIRONMENTAL EFFECTS
The mechanisms reviewed above have, for the most part, resulted in increases in
the macroscopic work of fracture. However, as in the case of the segregation of
species to certain interfaces, there are factors that can cause dramatic decreases in
the energy required to separate an interface. In particular, the environment can alter
the overall strength and reliability of an interface. A number of authors have noted
that certain interfaces, especially those where one of the materials composing the
bimaterial interface is SiO2, exhibit debond growth at applied driving energies
less than that required for critical fracture when the interface is delaminated in
a corrosive environment (e.g., 16, 77, 78). These authors have noted that this
behavior is similar to stress corrosion cracking in bulk glass systems. A large body
of work has focused on determining the mechanisms that control stress corrosion
in glasses, and detailed models have been developed to describe the controlling
processes (79–87). For simplicity, the discussion and examples herein are limited
4 Jun 2003 12:24 AR AR189-MR33-02.tex AR189-MR33-02.sgm LaTeX2e(2002/01/18) P1: IKH

INTERFACE FRACTURE 45

Figure 11 Effect of roughness of the macroscopic work of fracture for the


TiN/SiO2 interface. Lines represent the expected increase due to different mechanisms.
Roughness-induced shielding most closely matches the measured changes in interface
debond energy (48, 76).

to the effects of water on stress corrosion, but it should be noted that other corrosive
species (such as alcohols) can cause the same effect both in bulk glasses (e.g., 88)
and at interfaces (e.g., 77).
In bulk glasses, the environmental process leading to crack advance is a three-
step concerted reaction involving a water molecule and the strained Si-O bonds at
the crack tip (85–89). Plots of crack growth velocity versus crack driving energy
typically result in a sigmoidal shape (Figure 12), where changes in slope indicate
a change in the rate-limiting step for crack advance (e.g., 80, 82, 84). Region III is
dominated by critical fracture events and represents the interface fracture energy
in the absence of a corrosive species.
Region II, or the plateau region, is insensitive to crack driving force and the
rate-limiting step is found to be controlled by the diffusion of a reactive species to
the crack tip, so that the crack velocity is given by (80, 82, 84):
64 G a3o PH2 O
ν= ¡1¢ 1 , 12.
3π Eao ln ao x(2πmkT) 2
4 Jun 2003 12:24 AR AR189-MR33-02.tex AR189-MR33-02.sgm LaTeX2e(2002/01/18) P1: IKH

46 LANE

Figure 12 Schematic representing typical velocity versus driving force (V-G curve)
behavior for bulk glasses. The schematic shows the four distinct regions that correspond
to distinct rate-limiting steps for crack advance (79, 82, 84).

where E is Young’s modulus, l is the mean free path of the water vapor, m is the mass
of the diffusing species, P is the partial pressure of water vapor, x is the num-
ber of water molecules absorbed per bond, ao is the bond spacing, and G is the
mode I driving energy. The plateau velocity in Equation 12 depends linearly on
the partial pressure of water vapor, and an example of this occurring at interfaces
can be found in Figure 13 (48, 78).
As the crack velocity slows, the diffusion of water to the crack tip is greater than
the rate of crack advance, and the rate-limiting step switches from one controlled

Figure 13 (a) V-G curves for the Ta/SiO2 interface showing the shift in Region II with
partial pressure of water. (b) Liner relationship between the average Region II velocity and
partial pressure of water vapor γ (48, 78).
4 Jun 2003 12:24 AR AR189-MR33-02.tex AR189-MR33-02.sgm LaTeX2e(2002/01/18) P1: IKH

INTERFACE FRACTURE 47

by diffusion to one controlled by a stress-dependent chemical reaction occurring


at the debond tip (81, 82, 84). For the reaction given by
Si − O + H2 O → Si − (H2 O) − O, 13.

the fracture resistance is found to be (84)

0o = R = µSi-O∗ − µSi-O − µoH2 O − RTlnPH2 O , 14.

where µSi-O∗ , µSi-O, and µH2 O − RTlnPH2 O are the chemical potentials of the reactive
complex (Si-(H2O)-O), the crack tip bond, and the corrosive species, respectively.
The crack velocity in this region can be found by (84)
· ¸
G − N[α(T) − RTlnPH2 O ]
v = vo Sinh , 15.
η
where N is the number of bonds per unit area, T is the absolute temperature, PH2O
is the partial pressure of water vapor, α(T) represents the chemical potential of
the Si–O bond and the reactive complex, G is the applied driving force for crack
propagation, and η and ν o are parameters related to the bonds being broken. The
numerator of the Sinh function in Equation 15 implies a corrosion threshold below
which crack growth stops. Equation 15 also indicates that both the position of the
threshold driving force and the position of Region I will shift as the logarithm of
the partial pressure of water vapor. This behavior has been observed for different
interfaces with a representative example given in Figure 14 (48, 78). Note that
Figure 14b represents the effect of both temperature and humidity and that tem-
perature has only a weak effect on the position of the corrosion threshold. This
is consistent with Equation 15 in that over the range measured, the temperature
changed by a maximum of 29%, while the natural log of the partial pressure of
water vapor changed by 150%.
In some systems, an alternative threshold regime has been observed (85) in
which the threshold occurs when the corrosive molecule can no longer reach the
debond tip because it is sterically hindered from doing so. In this case, the thresh-
old depends only on the crack opening displacement and the size of the corrosive
species and only very weakly on the environment. The data in Figure 13 have
been replotted in Figure 15 with the onset of the threshold region expanded, which
illustrates this phenomenon. As mentioned above, Region II behavior consistent
with that expected in bulk specimens is observed for the interface. However, there
is a transition from Region II directly to a threshold as GI is lowered, with the
position only weakly dependent on the test conditions. This transition to a steric
hindrance threshold was shown to be consistent with the formation of a Ta oxide,
the thickness of which depends on the time of exposure (determined by the plateau
velocity) and the test temperature, which blocked the water molecule from reaching
the debond tip by reducing the opening behind the debond tip according to (78)
√ r
2 GI 8y dH O
uop = ao + √ − tox − 2 , 16.
E π 2
4 Jun 2003 12:24 AR AR189-MR33-02.tex AR189-MR33-02.sgm LaTeX2e(2002/01/18) P1: IKH

48 LANE

Figure 14 (a) Region I and threshold driving force dependence on the environment.
(b) Change in threshold driving force as a function of temperature and partial pressure
of water vapor, indicating that temperature has only a small effect (48, 78).

where the first two terms represent the opening expected for a linear elastic solid,
where ao is the bond spacing, GI is the mode I driving energy, and y is the position
behind the debond tip. This opening displacement is reduced by the length of a
hydroxyl (-OH) group, which accounts for the rupture process and is represented
by d H2 O /2 and the thickness of the oxide, tox. The onset of the steric hindrance
4 Jun 2003 12:24 AR AR189-MR33-02.tex AR189-MR33-02.sgm LaTeX2e(2002/01/18) P1: IKH

INTERFACE FRACTURE 49

Figure 15 Transition from Region II behavior directly


to a steric threshold regime that is only slightly dependent
on the environment (48, 78).

threshold is found by solving Equation 16 for GI. A plot of the change in measured
threshold driving energy found in Figure 15 and that predicted by Equation 16 can
be found in Figure 16.
Finally, Equation 14 indicates that the effective work of adhesion can be altered
by changing the environment in which an adhesion test is conducted. By doing
this, the relationship between the work of adhesion and plastic energy dissipation
can be explored. Figure 17 (48, 78) illustrates this by plotting the change in work of
adhesion, predicted by Equation 14, for the TaN/SiO2 interface versus the macro-
scopic work of fracture. The result is a straight line with a slope of ∼20, which
indicates that for every 1 J/m2 the work of adhesion is changed as the macroscopic
work of fracture changes by a factor of 20. This is consistent with predictions of
interface fracture models (69).

RELATIONSHIP BETWEEN INTERFACIAL ADHESION


AND OTHER INTERFACIAL PHENOMENON
It can be expected that relationships exist between interfacial adhesion and other
interfacial phenomena that are controlled by the same mechanisms such as the
strength of interfacial bonds. Consider the particular example of electromigration
in Cu conductors, a common concern in the microelectronics industry [for a discus-
sion of electromigration see for example (90)]. Briefly, electromigration is the dif-
fusion of atoms in the direction of electron flow. In Cu conductors, the diffusion path
is found along a particular interface (91) and the adhesion of that interface has been
4 Jun 2003 12:24 AR AR189-MR33-02.tex AR189-MR33-02.sgm LaTeX2e(2002/01/18) P1: IKH

Figure 16 Predicted change in steric hindrance threshold (solid line) versus that
measured in Figure 15. The predicted change is based on the growth of an oxide layer
immediately following fracture of the Ta/SiO2 interface (48, 78).

Figure 17 Relationship between the macroscopic work of fracture and the work of
adhesion for the TaN/SiO2 interface with a 16.4-µm Cu layer adjacent to the interface.
As the effective work of adhesion is decreased (by increasing the moisture content in
the testing environment) the plastic energy dissipation in the ductile Cu layer is reduced
(48, 78).

50
4 Jun 2003 12:24 AR AR189-MR33-02.tex AR189-MR33-02.sgm LaTeX2e(2002/01/18) P1: IKH

INTERFACE FRACTURE 51

Figure 18 Relationship between interface adhesion and electromigration activation


energy. The liner relationship has been predicted by simple bond-breaking models and
indicates that interface adhesion measurements may be used as screening devices for
other processes that are dependent on interfacial strength (93).

speculated to control the diffusion (91, 92). Recently (93), work has focused on
deriving a relationship between the adhesion of the fast diffusion path and the ele-
ctromigration activation energy, the results of which are summarized in Figure 18.
The linear relationship between the work of adhesion and electromigration ac-
tivation energy has been predicted (93, 94) and should not come as a surprise,
since both diffusion and fracture involve, to one extent or the other, the breaking
of bonds. There are two implications of this. The most obvious is that interfa-
cial adhesion measurements can be used as a screening process to address other
interfacial phenomena that may require laborious and costly sample preparation
(as in the case of electromigration samples), so long as a link between the two
measurements can be established. The second perhaps more subtle implication is
that bond character controls these interfacial phenomena and that the measure of
the bond character would give the ultimate insight into the interfacial phenomena.
For example, consider the work of Muller and coworkers (95), who investigated
the effects of B segregation to Ni-Al grain boundaries with electron-energy loss
spectroscopy (EELS). Muller et al. showed that B segregation altered the electronic
structure of Ni in the grain boundaries of a Ni3Al intermetallic and correlated that
change to a change in fracture properties of the grain boundary. Later, Rosenberg,
working with Muller (96), demonstrated that EELS spectra revealed changes in
4 Jun 2003 12:24 AR AR189-MR33-02.tex AR189-MR33-02.sgm LaTeX2e(2002/01/18) P1: IKH

52 LANE

the electronic structure of Cu at Al grain boundaries in Al-Cu films and correlated


that change with results from diffusion experiments. These results are tantalizing
in that they point to a relationship between apparently different interfacial phe-
nomena (such as adhesion and electromigration) and indicate that insight into the
relationship between them can be elucidated with investigation of the electronic
structure of the interfacial bonds.

SUMMARY

Through models developed to describe the controlling mechanisms of interfa-


cial adhesion, researchers can accurately describe a number of disparate energy-
absorbing processes. The utility of these models lies in the basic trends they predict,
which can serve as a basis or starting point for constructing robust interfaces. Al-
ternatively, these models allow for the deconstruction of the macroscopic work
of fracture into its contributing energies so that only those energy-absorbing pro-
cesses that are important for a specific application can be examined. Of all the
energy-absorbing processes, it is the work of adhesion that yields the most influ-
ence, as all other processes scale with it. Because the work of adhesion is a direct
measure of the bond strength across the interface, electron spectroscopies such
as EELS, which measure electron configurations of atoms, may provide a link
between electron sharing or bonding and the work of adhesion.

The Annual Review of Materials Research is online at


http://matsci.annualreviews.org

LITERATURE CITED
1. Jokl ML, Vitek V, McMahon CJ. 1980. Acta 10. Cao HC, Evans AG. 1989. Mech. Mater.
Metall. 38:1479 7:295–304
2. Evans AG, Rühle M, Dalgleish BJ, Char- 11. Wang JS, Suo Z. 1990. Acta Metall. Mater.
alambides PG. 1990. Metall. Trans. A 38:1279–90
21A:2419–29 12. Jensen HM, Thouless MD. 1993. Int. J.
3. Cannon RM, Fisher RM, Evans AG. 1986. Solids Struct. 30:779–95
MRS Proc. 54:799–804 13. Hutchinson JW, Suo Z. 1992. Adv. Appl.
4. Cannon RM. 1984. Adv. Ceram. 10:818 Mech. 29:63–187
5. Shih CF. 1991. Mater. Sci. Eng. A 143:77– 14. Dundurs JJ. 1969. Appl. Mech. 36:650–
90 52
6. Ozdil F, Carlsson LA. 1992. Eng. Fract. 15. Fleck NA, Hutchinson JW, Suo Z. 1991.
Mech. 41:645–58 Int. J. Solid Struct. 27:1683–703
7. Tvergaard V, Hutchinson JW. 1994. Philos. 16. Xu G, He M-Y, Clarke DR. 1999. Acta
Mag. A 70:641–56 Mater. 47:4131–41
8. Evans AG, Hutchinson JW. 1989. Acta 17. Dauskardt RH, Lane MW, Ma Q, Krishna
Metall. 37:909–15 N. 1998. Eng. Fract. Mech. 61:141–62
9. Liechti KM, Chai YS. 1992. Appl. Mech. 18. Volinsky AA, Moody NR, Gerberich WW.
59:295–304 2002. Acta Mater. 50:441–66
4 Jun 2003 12:24 AR AR189-MR33-02.tex AR189-MR33-02.sgm LaTeX2e(2002/01/18) P1: IKH

INTERFACE FRACTURE 53

19. Marshall DB, Evans AG. 1984. J. Appl. Ritchie RO. 1997. Mech. Mater. 25:291–
Phys. 56:2632–38 308
20. Turner MR, Evans AG. 1996. Acta Mater. 43. Mittal KL. 1978. ASTM Spec. Tech. Publ.
44:863–71 Vol. 640
21. Kriese MD, Gerberich WW. 1999. J. Mater. 44. Rühle M, Evans AG. 1989. Mater. Sci. Eng.
Res. 14:3007–18 A 107:187–97
22. Kriese MD, Moody NR, Gerberich WW. 45. Griffith AA. 1920. Philos. Trans. R. Soc.
1998. Acta Mater. 18:6623–30 London Ser. A 221:163–98
23. Kriese MD, Moody NR, Gerberich WW. 46. Rice JR, Wang J-S. 1989. Mater. Sci. Eng.
1997. MRS Proc. 473:39–49 A 107:23–40
24. Volinsky AA, Tymiak NI, Kriese MD, Ger- 47. Lane MW, Dauskardt RH, Krishna N,
berich WW, Hutchinson JW. 1998. MRS Hashim I. 2000. J. Mater. Res. 15:203–
Proc. 539:277–90 11
25. Bagchi A, Lucas GE, Suo Z, Evans AG. 48. Lane MW. 2000. Adhesion in multi-layer
1994. J. Mater. Res. 9:1734–41 thin film systems: chemistry, morphology,
26. Xu G, Ragan DD, Clarke DR, He MY. plasticity and environmental effects. PhD
1997. MRS Proc. 458:465 thesis. Stanford Univ. 150 pp.
27. Xu G, Mates TE, Clarke DR, Ma Q, Fuji- 49. Lipkin DM, Clarke DR, Evans AG. 1998.
moto H. 1999. MRS Proc. 563:257–62 Acta Mater. 46:4835–50
28. Suo Z, Hutchinson JW. 1989. Mater. Sci. 50. Korn D, Elssner G, Cannon RM, Rühle M.
Eng. A 107:135–43 2002. Acta Mater. 50:3881–901
29. Kanninen MF. 1973. Int. J. Fract. 9:83–92 51. Cannon RM, Korn D, Elssner G, Rühle M.
30. Atkinson C, Smelser R, Sanchez J. 1982. 2002. Acta Mater. 50:3903–25
Int. J. Fract. 18:279 52. Zhuk AV, Evans AG, Hutchinson JW,
31. Charalambides G, Lund J, Evans AG, Whitesides GM. 1998. J. Mater. Res. 13:
McMeeking RM. 1989. J. Appl. Mech. 3555–64
111:77–82 53. Lin J. 1999. Adhesion at the PMMA/Si in-
32. Fernlund G, Spelt JK. 1994. J. Compos. terface. MS thesis. Stanford Univ. 18 pp.
Tech. Res. 16:234–43 54. Lane MW. 2002. Appl. Phys. Lett. Submit-
33. Fernlund G, Spelt JK. 1991. Int. J. Adhes. ted
Adhes. 11:213–20 55. Cook RF, Liniger EG. 1993. J. Am. Ceram.
34. Fernlund G, Spelt JK. 1991. Int. J. Adhes. Soc. 76:1096–105
Adhes. 11:221–27 56. Wei Y, Hutchinson JW. 1997. J. Mech.
35. Fernlund G, Spelt JK. 1994. Comp. Sci. Phys. Solids 45:1253–73
Tech. 50:441–49 57. Tvergaard V. 1999. J. Mech. Phys. Solids
36. Wylde JW, Spelt JK. 1998. Int. J. Adhes. 47:1095–112
Adhes. 18:237–46 58. Klein PA, Foulk JW, Chen EP, Wimmer SA,
37. Swandener JG, Liechti KM, de Lozanne Gao HJ. 2001. Theor. Appl. Fract. Mech.
AL. 1999. J. Mech. Phys. Solids 47:223– 37:99–166
58 59. Klein P, Gao H. 1998. Eng. Fract. Mech.
38. Liang Y-M, Liechti KM. 1995. Int. J. Solids 61:21–48
Struct. 32:957–78 60. Shih CF, Asaro RJ. 1990. Int. J. Fract.
39. Ma Q. 1997. J. Mater. Res. 12:840–45 42:101–16
40. Snodgrass JM, Dauskardt RH. 2001. MRS 61. Zhang P, Klein P, Huang Y, Gao H, Wu PD.
Proc. 612:1–6 2002. CMES 3:263–77
41. Lane MW. 2002. Interface adhesion. Pre- 62. Tvergaard V, Hutchinson JW. 1996. Int. J.
sented at ASME Meeting, Blacksburg, VA Solids Struct. 33:3297–308
42. Becker TL Jr, McNaney JM, Cannon RM, 63. Klein PA, Gao H, Vainchtein A, Fujimoto
4 Jun 2003 12:24 AR AR189-MR33-02.tex AR189-MR33-02.sgm LaTeX2e(2002/01/18) P1: IKH

54 LANE

H, Lee J, Ma Q. 2000. MRS Proc. 594:371– 80. Wiederhorn SM. 1967. J. Am. Ceram. Soc.
77 50:407–13
64. Suo Z, Shih CF, Varias AG. 1993. Acta Met- 81. Wiederhorn SM, Bolz LH. 1970. J. Am. Ce-
all. Mater. 41:1551–57 ram. Soc. 53:543–48
65. Wei Y, Hutchinson JW. 1999. Int. J. Fract. 82. Lawn BR. 1975. J. Mater. Sci. 18:469–80
95:1–17 83. Lawn BR. 1974. Mater. Sci. Eng. 13:277–
66. He MY, Evans AG, Hutchinson JW. 1996. 83
Acta Mater. 44:2963–71 84. Cook RF, Liniger EG. 1993. J. Am. Ceram.
67. Needleman A. 1987. J. Appl. Mech. 54: Soc. 76:1096–105
525–31 85. Michalske TA, Frieman SW. 1983. J. Am.
68. Beltz GE, Rice JR, Shih CF, Xia L. 1996. Ceram. Soc. 66:284–88
Acta Mater. 44:3943–54 86. Michalske TA, Bunker B. 1987. Sci. Am.
69. Lane MW, Dauskardt RH, Vainchtein A, 257:78–85
Gao H. 2000. J. Mater. Res. 15:2758–69 87. Michalske TA, Bunker B. 1987. J. Am. Ce-
70. Cotterell B, Rice JR. 1980. Int. J. Fract. ram. Soc. 70:780–84
16:155–69 88. Wiederhorn SM, Freiman SW, Fuller ER,
71. Williams JG, Ewing PD. 1972. Int. J. Fract. Thomson R. 1982. J. Mater. Sci. 17:3460–
Mech. 8:441–46 78
72. Gell M, Smith E. 1967. Acta Metall. 15: 89. Mencik J. 1992. Strength and Fracture of
253–58 Glass and Ceramics. New York: Elsevier
73. Lawn B. 1993. Fracture of Brittle Solids. 90. Rosenberg R, Edelstein DC, Hu C-K, Rod-
New York: Cambridge Univ. Press. 2nd ed. bell KP. 2000. Annu. Rev. Mater. Sci. 30:
74. Ma Q, Fujimoto H, Flinn P, Jain V, Adivi- 229–62
Rizi F, Dauskardt RH. 1995. MRS Proc. 91. Hu C-K, Rosenberg R, Lee KL. 1999. Appl.
391:91–96 Phys. Lett. 74:2945
75. Clarke DR. 1997. MRS Proc. 473:27– 92. Lloyd JR, Clement JJ. 1995. Thin Solid
38 Films 262:135–41
76. Lane MW, Ni W, Dauskardt RH, Ma Q, 93. Lane MW, Liniger EG, Lloyd JR. 2003. J.
Fujimoto H, Krishna N. 1997. MRS Proc. Appl. Phys. 93:In press
473:357–62 94. Zener C. 1952. In Imperfections in Nearly
77. Card JC, Cannon RM, Dauskardt RH, Perfect Crystals. ed. W Shockley, p. 289.
Ritchie RO. 1993. MRS Proc. 314:109– New York: Wiley
16 95. Muller DA, Subramanian S, Batson PE, Sil-
78. Lane MW, Snodgrass JM, Dauskardt RH. cox J, Sass SL. 1996. Acta Mater. 44:1637–
2001. Microelectron. Reliability 41:1615– 45
24 96. Rosenberg R, Batson PE, Bruley J, Splin-
79. Wiederhorn SM. 1968. Int. J. Fract. Mech. ter J, Muller DA. 1998. AIP Conf. Proc.
4:171–77 418:127–34

Você também pode gostar