You are on page 1of 19

International Journal of Fracture 90: 133151, 1998.

1998 Kluwer Academic Publishers. Printed in the Netherlands.

Energy balance in dynamic fracture, investigated by a


potential drop technique

J.A. HAUCH and M.P. MARDER


Center for Nonlinear Dynamics and Department of Physics, University of Texas at Austin, Austin, Texas 78712,
U.S.A.

Received 23 January 1997; accepted in revised form 8 January 1998

Abstract. A puzzling question in dynamic fracture has been why cracks in amorphous brittle materials always
travel at velocities smaller than the Rayleigh wave speed. The answer is that the energy per length needed for the
crack to propagate depends strongly on velocity. As the energy flux to the crack tip increases, the crack chooses
new modes of dissipation such as micro-cracking and the creation of subsurface damage zones to dissipate this
energy. In this paper we use a potential drop technique to measure length and velocity of a crack with high
spatial precision and time resolution so as to investigate the modes of dissipation in Homalite-100 and make
qualitative comparisons with PMMA. The technique is capable of resolving crack initiation, run, and arrest. Using
this technique we search for a forbidden band of velocities in PMMA, Homalite-100, and glass, and we show
that no such velocity gap exists in these amorphous materials at room temperature.

Key words: Energy balance, dissipation, fracture energy, velocity gap, potential drop technique.

1. Introduction

Ever since Griffith broke glass rods in his laboratory in England and realized that flaw growth
is only possible if the energy released by crack advance is larger than the energy needed to
create the surface (Griffith, 1920), energy and how it is dissipated has been at the heart of
quantitative fracture mechanics. In 1947 Mott realized that the inclusion of a kinetic energy
term in Griffiths framework can extend this energy approach to encompass dynamic fracture
(Mott, 1947). He found that the velocity of a crack should asymptotically approach a terminal
velocity, which Stroh proposed should be the Rayleigh wave speed in the material (Stroh,
1957). In fact this result was already implicit in the results of Yoffe (Yoffe, 1951). As corrected
by Dulaney and Brace in 1960 (Dulaney and Brace, 1960) Motts scaling argument remained
essentially unscathed, despite tremendous increases in the mathematical sophistication of dy-
namic fracture mechanics over the next years (Freund, 1990). However there was a difficulty.
Cracks in brittle amorphous materials never were observed to reach the Rayleigh wave speed
(Kobayashi et al., 1974; Irwin et al., 1979).
This difficulty was always more apparent than real. On the one hand, cracks on the cleavage
planes of brittle crystals, or along weak interfaces did achieve velocities close to the Rayleigh
wave speed (Gilman et al., 1958; Hull and Beardmore, 1966; Cotterell, 1965; Field, 1971).
On the other, no theory of dynamic fracture was able to make any predictions about the
velocity of a crack without making a presumption about the energy needed per unit length
to propagate a crack. Most dynamical equations assumed this quantity to be a constant (Hall,
1953). However, the experimental results can easily be explained by dropping this assumption
and allowing fracture energy to increase with velocity. For a long time it has been known that
134 J.A. Hauch and Marder

cracks in brittle polymers at high stress intensity factors attempt to branch, with the branches
subsequently dying out leaving behind a series of short micro-cracks at an angle to the main
crack (Irwin et al., 1979; Kobayashi and Mall, 1978; Kobayashi and Dally, 1977; Dally, 1979;
Ravi-Chandar and Knauss, 1984a). In the brittle plastics PMMA and Homalite-100 Doyle
(1983) and Ravi-Chandar and Knauss (1984) showed that the increase in fracture energy
was connected to the generation of micro-cracks underneath the fracture surface. The central
question in dynamic fracture mechanics therefore was reduced to determining why energy
consumption should rise so abruptly past a critical threshold, or putting it more properly, why
when energy flux to the crack tip exceeded a threshold, the mean velocity stopped increasing.
A tentative answer to this question has been given in the context of lattice calculations.
In analytically solvable models as well as extensive computer simulations, it has been shown
that when energy flux to a crack tip passes a critical threshold, the tip becomes unstable,
sometimes to micro-cracks and sometimes to dislocations (Liu and Marder, 1991; Marder and
Liu, 1993; Marder, 1993; Marder and Gross, 1994; Abraham et al., 1994; Zhou et al., 1994).
Some of the images produced by these simulations bear great visual resemblance to images
of PMMA past its threshold of instability (Sharon et al., 1995; Sharon et al., 1996). The
atomic scale calculations make additional predictions as well. The most striking is that there
should be a range of velocities, roughly between 0 percent and 20 percent of the Rayleigh
wave speed, at which steady crack motion would be impossible (Marder, 1993; Marder et
al., 1994), a process also known as lattice trapping (Thomson et al., 1971; Sinclair and Lawn,
1972; Gumbsch, 1995). All the calculations are really appropriate for comparison with crystals
at low temperature, not glass at one third of its melting temperature. Still one might easily
wonder if in brittle amorphous solids such a velocity gap exists.
Our purposes in this paper are threefold. First we intend to relate details of the dynamic
potential drop method we have employed for several years to investigate crack dynamics. The
method is comparatively simple and inexpensive, and our discussion may enable other groups
to utilize it. Second we investigate whether micro-cracks in Homalite-100 provide the same
qualitative explanation for the rise in fracture energy that Sharon et al. have demonstrated in
PMMA. Finally we ask whether a velocity gap exists in glass, PMMA, and Homalite. Despite
previous misleading indications the answer is no.

2. Experimental technique

In dynamic fracture mechanics crucial data allowing quantitative and qualitative comparison
with theory are the length and the velocity of the crack as a function of time or each other, and
the stress state of the sample. These quantities cannot be obtained through measurement with
a single localized probe. Several measurement techniques have been developed to measure
these quantities. The techniques are variations on four major themes.
(1) One is high-speed photography, from which measurement of crack velocities have been
obtained for over 50 years (Schardin et al., 1955; Dally, 1979; Ravi-Chandar and Knauss,
1984a, 1984b, 1984c, 1984d; Knauss and Ravi-Chandar, 1985, Field, 1971). This technique
has two limitations in measuring fast fracture. First, cameras cannot be triggered to capture
crack motion unless the time of initiation is known with some certainty (Wenner and Rogers,
1977). This fact dictates rather violent loading conditions for such experiments, since these
can be guaranteed to produce crack motion in a short window of time. Second, it has been
impossible to gather more than 128 frames in a single experiment (Ravi-Chandar and Knauss,
Energy balance in dynamic fracture 135

1982), and therefore the time series for crack motion cannot be known in great detail. An
advantage of this technique however is that in a variety of birefringent materials it can also
yield information about the dynamic stress intensity factor at the crack tip in addition to the
crack length.
(2) A second is the resistive grid technique, in which a series of resistive wires is laid down
on the sample, and is broken by a running crack (Kobayashi et al., 1974; Wenner and Rogers,
1977; Cotterell, 1965). Sample preparation for this method is usually difficult and resolution
is limited by the experimenters ability to lay down thin wires and interface with them.
(3) A third method that has also been applied widely is acoustic modulation of the crack
tip. In this type of experiment the crack tip is deflected by ultra-sound waves introduced into
the specimen, leaving behind a set of periodic markings that can then be measured and reflect
the velocity of the crack tip (Field, 1971; Kerkhof, 1973). This type of measurement yields
length and velocity data; however, it is intrinsically intrusive and may modify the dynamics
of the crack tip, particularly in the presence of instabilities. Another disadvantage is that the
technique is limited to materials that produce smooth fracture surfaces, and have low acoustic
attenuation.
(4) Potential drop techniques measure crack length through changes in the resistance pro-
duced by crack propagation in a specimen. The method has widely been used to measure
quasi-static crack growth for conductive specimens (Wenner and Rogers, 1977; Buck, 1989).
It can be adapted to measure the surface velocity in insulators by applying conductive coatings
(Stalder et al., 1983; Bguelin et al., 1983; Stalder and Kausch, 1985; Stalder et al., 1989;
Cudr-Mauroux et al., 1991; Wang and Kim, 1993). The resolution of this technique is limited
only by deviation of the crack from the straight cut assumed in analysis, and by the speed of the
acquisition equipment. Crack velocity can therefore be obtained at higher rates, with greater
accuracy, and greater density of measurements than other techniques. In addition, the method
can be applied to a wide range of materials and fracture processes ranging from polymers to
single crystals, and from fast fracture and crack arrest to interface fracture in virtually every
specimen geometry.
In this paper we use a variation of the potential drop technique that has been developed at the
Center for Nonlinear Dynamics. The technique yields measurements of the crack length with
a resolution of 0.2 mm and the crack velocity with a resolution of 10 m/s or better at a temporal
resolution of 5 MHz. With this method data sets of 64000 points for both the length and the
velocity are obtained on both sides of the specimen simultaneously, allowing full resolution
of initiation, run and arrest for fracture processes lasting up to 3.2 ms. We use the technique
to investigate the dynamics of cracks in PMMA, Homalite-100 and glass.

2.1. C OATING

The main ingredient of potential drop techniques is a conductor that is ruptured during the
fracture process. As the fracture proceeds, the resistance of the conductor changes, and pro-
vides information on the length and velocity of the crack. The conductor may be the fracture
sample itself, or if the specimen is an insulator, it can be a thin metallic coating that is applied
to the fracture sample. All experiments described in this paper concern insulating samples
coated with aluminum, since aluminum is brittle and a good conductor. The coating is evapo-
rated onto the samples in a vacuum of 106 Torr. To assure that the coating does not affect the
physical properties of the sample its thickness must be very small in comparison to the sample
thickness. In our experiments the coating thickness is varied with sample size but is between
136 J.A. Hauch and Marder

Figure 1. The figure shows the coating geometry and electrode placement on the fracture sample. Plotted above the
plate is the current density in the coating obtained by solving Laplaces equation. The graph shows the resistance
as a function of crack length, simulation and calibration. The inlay is an enlargement of the calibration data. It
demonstrates that there are no small scale variations within the resolution of the measurement.

2050 nm in all cases. The resistance as a function of crack length is sensitive to the geometry
of the coating, and placement of the electrodes on the sample. Coating geometry and electrode
placement used in our experiments are shown in Figure 1. This particular configuration was
chosen for the linearity of resistance as a function of crack length, which can be obtained
by solving Laplaces equation subject to the appropriate boundary conditions. A surface plot
of the current density in the coating is depicted in Figure 1. To ensure good agreement with
simulation the functional form of the resistance was checked in PMMA by cutting the coating
and measuring the resistance. Figure 1 compares the two measurements, showing only slight
discrepancies in the beginning and the end that can be attributed to the evaporation process
by which the aluminum was deposited on the sample. The evaporation geometry results in a
slightly thicker coating in the middle of the sample. To alleviate this problem it is possible to
take the samples after the experiment and measure the functional form of the resistance on a
sample by sample basis in detail. This was not done in our experiments since the variations
were found to be less than the resolution of the digitizers used. The average resistance gradient
however was measured on each sample by bridging the crack with a piece of copper in order
to account for variations in the average thickness of the coating between samples, which could
be as high as 15 percent.
To gain confidence in the method several issues need to be addressed. When the samples
are loaded there will be a change in the cross sectional area of the coating that will affect its
resistance. The change in area is determined by sample strain and the Poisson ratio of the
coating material, and is given by: A/A = 2 ' 0.3450.12 = 3.45105 . This will result
in a maximum shift in resistance of: R = A/A Rmax ' 50 3.45 105 ' 2 103 ,
which is an order of magnitude smaller than the resolution of the instrumentation.
Another concern is electrical discharge across the crack. The potential drop between the
electrodes increases with the length of the crack, but never exceeds 3 V over the useful range
of the measurements. The largest possible potential drop is then across the crack at the position
of the electrodes. To exceed a critical electric field of 10 6 V/m for sparking in air this entire
voltage drop must occur over a distance of 3 m. The separation of the crack faces at the
position of the electrodes will be on the order of the total strain of the plate or about 50 m
in the polymers, which makes discharge impossible. In glass, however, the separation of the
Energy balance in dynamic fracture 137

crack faces is much smaller, only about 23 m, and electrical discharge has been observed in
our experiments. The process lasts about 1 s and is clearly distinguishable from the fracture
dynamics, since it is characterized by an apparent jump to a negative velocity followed by a
rapid apparent jump to a large positive velocity that quickly decays.
Small scale variations of the coating thickness could also affect the data. The method will
be most sensitive to variations in the coating at the position of the peak in the current density
in front of the crack tip (Figure 1). The full width at half maximum of this peak should thus
determine a length-scale on which variations in the coating thickness will cause problems.
From the simulations the full width at half maximum is found to be greater than 1 percent of
the plate length which is on the order of millimeters. The length-scale on which the coating
thickness is expected to vary due to the vapor deposition process can be estimated by the vapor
droplet size of the aluminum which is on the order of 1 nm. No statistical variations can be ex-
pected for length-scales on the order of 1000 droplets, about 1 m. Thus thickness variations
in the coating pose no problem since the method is inherently an averaging process involving
a large area of the coating. In addition it can be seen from the inset of Figure 1 that there are no
measurable small scale variations that are within the resolution of the measurement method.
Generation of static charge on the crack faces could also interfere with the measurements.
We can estimate an upper bound for the currents due to charge generation if we assume one
electron per bond will end up on one of the crack faces, and a bond density of 5 1012 bonds
per square millimeter. A crack traveling at 300 m/s a 3.2 mm thick plate will break bonds at a
rate of 5 1018 bonds/s. If all the electrons end up on the same crack face this could lead to
a current of 800 mA. To preserve charge neutrality, this charge might generate an additional
current loop around the crack tip, but due to the principle of superposition it could have no
effect on the measurements at the position of the electrodes.

2.2. I NSTRUMENTATION

To instrument the samples electrodes are attached on either side of the seed crack using a
conductive paint that is made from a suspension of silver particles. The electrodes are part
of a Wheatstone bridge powered by batteries for noise reduction. The output of the bridge is
digitized at a rate of 10 MHz by a Burr-Brown ZDP1002 12 bit Analog to Digital converter. In
addition the output of the bridge is also fed into an analog differentiator that outputs a voltage
proportional to the time rate of change of the input signal, and is then digitized at 20 MHz by
a Markenrich WaagII 8 bit Analog to Digital converter. Details of the electronics are shown in
Figure 2.
The bridge voltage time series can be inverted to find the length of the plate using the
expression
 
Vbat (R1 + Ra )
R P l = Rb 1 , (1)
Ra Vbat VBr (R1 + Ra )

where RP l is the plate resistance, Vbat is the input voltage to the bridge, and VBr is the output
voltage of the bridge. The resistances R1 , Ra , and Rb are the other three resistances in the
bridge as defined in Figure 2. Using a lookup table generated by the solution of Laplaces
equation, or by direct measurement on the sample RP l can then be converted to crack length. A
simpler procedure is to consider only the data in the linear resistance regime and approximate
138 J.A. Hauch and Marder

Figure 2. The diagram shows the details of the instrumentation used in the experiments. The values of Ra = 500 
and Rb = 250  were chosen to minimize the power dissipated in the plate. R1 is chosen individually in each
experiment to match the initial bridge output to the range of the digitizers, and ranges from 30 to 70.

RP l by a linear function with an offset R0 due to contact resistance and a slope determined by
the derivative of RP l

dRP l
RP l = R0 + l, (2)
dl
in which case an expression for the length l of the crack is obtained
   
RP l R0 dl Vbat (R1 + Ra )
l= = Rb 1 R0 . (3)
dRP l
dl
dRP l Ra Vbat VBr (R1 + Ra )

The time rate of change of the output of the Wheatstone bridge, VBr is determined by

dVBr Rb dRP l Rb dRP l dl


= Vbat = Vbat . (4)
dt (Rb + RP l ) dt
2 (Rb + RP l ) dl dt

Crack velocity can now be extracted using the calibrated response of the analog differentiators

dVBr
Vdif = K + Vo , (5)
dt
where Vdif is the differentiator output voltage and Vo is an arbitrary offset voltage, and com-
bining it with the voltage time series from the Wheatstone bridge to obtain

dl (Vo Vdif) (Rb + RP l )2 dl


= . (6)
dt KVbat Rb dRP l
Energy balance in dynamic fracture 139

Figure 3. Crack length and velocity data froma full thin strip experiment in Homalite-100. The stress intensity
factor for this experiment was KI = 0.52 MPa m which is in region where the crack velocity is highly sensitive
to KI . The linear resistance approximation of (2) was used to convert the data. The data clearly show that this
approximation starts to fail for crack lengths greater than 23 cm for the 28 cm plate, which is to be expected from
the calibration. In all later analysis data that falls into the nonlinear region is ignored.

Every quantity in this expression is known either from direct measurement or deduction. The
precision of the measurements is usually limited by internal noise of the electronics, but can
be manipulated by adjusting the input voltage of the bridge and the coating thickness which
determines the magnitude of dRP l /dl in (2). Typical resolution in the experiments performed
is 0.2mm for the crack length and 10 m/s for the crack velocity at a bandwidth of 5 MHz. A
full data set for both length and velocity for a thin strip experiment in Homalite-100 is shown
in Figure 3.
To ensure consistency in the experiments these measurements are done on both sides of
the sample simultaneously, resulting in two data sets for each experiment. The ZPD1002
has a 256 Kbyte buffer for two channels, resulting in a 64 K sample, 6.4 ms long data set at
12 bit resolution for each side of the plate. Similarly the WaagII has a 128 Kbyte buffer for 2
channels, yielding a 64 K sample, 3.2 ms long data set at 8 bits for each side. Synchronization
is achieved by triggering both cards simultaneously with a comparator when the outout voltage
of the Wheatstone bridge passes through 0 V, coinciding approximately with the middle of the
plate. Both cards are set up such that they store pre-trigger information in half of their buffer
capacity, and post-trigger information in the other half. The setup yields information for the
entire fracture process in a typical experiment with duration between 13 ms.

2.3. T HIN STRIP LOADING CONFIGURATION

One set of experiments was done in a thin strip configuration with fixed displacement bound-
ary conditions identical to the loading used in (Fineberg et al., 1991, 1992; Gross et al., 1993;
Marder and Gross, 1995; Sharon et al., 1995, 1996). Thin strip geometry refers to experiments
with a large aspect ratio L/W (Figure 4, i.e. specimen that are narrow and long). The advantage
of this loading geometry is that it makes the task of determining energy flux to the crack tip
trivial. Once a fracture achieves steady-state conditions one has necessarily that the fracture
energy G and the stress intensity factor KI are determined by

KI2 2 E 2
G= = W= , (7)
E 2E 2W
140 J.A. Hauch and Marder

Figure 4. Loading configurations: (a) The thin strip configuration was used in constant fracture energy experi-
ments. The plate is loaded by displacing the edges by a constant distance . (b) The Double Cantilever Beam
configuration was used in crack arrest experiments. The plate is loaded by displacing the circular discs at a constant
rate . During fracture the conditions are assumed to be quasi-static.

where E is Youngs modulus, is the stress far from the crack tip and and W are defined
in Figure 4. Steady-state here means that the crack has come into dynamic equilibrium with
waves reflecting from the upper and lower boundaries, but is sufficiently far from beginning
and end of the sample. Sample dimensions in these experiments varied with an aspect ratio
L/W between 3.3 and 3.6 in all experiments, but most experiments were done on specimen
that measured 20 cm 5.5 cm with a thickness of 0.32 mm. Using these sample dimensions
and the two conditions for steady-state shows that the data for which the expression is valid
are acquired in the time interval from 60 s to 400 s. Since the fracture energy is fixed,
constant velocity experiments (see Figures 3 and 5) can be realized in this configuration. As a
result, observations can be performed on comparatively large areas of fracture surface created
at the same fracture energy. In early experiments the thin strip boundary conditions were
realized by attaching oil hardened precision ground steel strips to the edges of the samples
with Permabond 910 industrial grade adhesive. With this method in Homalite-100 the stress
variation along the length of the plate was less than 3 percent of the total stress. In later exper-
iments a clamp type jaw was used and the samples were glued directly to the jaws, resulting in
improved boundary conditions. The edges were then displaced in our home built screw type
tensile testing machine by a constant distance in a sequence of steps, such that the stress in
the sample increased by about 1 percent of the anticipated failure stress during each step. If
failure did not occur within one minute, the process was repeated. In a successful experiment
the failure occurred 1060 s after the stepping process so that effects due to the stepping of the
jaws could be neglected. In order for the boundary conditions to remain constant throughout
Energy balance in dynamic fracture 141

the experiment, jaw motion due to machine compliance during the fracture process must also
be kept to a minimum. The jaw motion can be estimated with the following expression

1 F t3
jaw (t) = . (8)
6 M tbreak

Where jaw is the displacement of the jaw, F is the total load applied, M is the jaw mass,
t is time and tbreak is the total duration of the fracture process. It should be noted that jaw
is independent of the compliance of the tensile testing machine, and solely depends on the
applied load, mass of the jaw and experiment duration. The total variation in stress due to
this effect is on the order of 2 percent, but it is proportional to the time cubed which makes
it negligible for the initial stages of the experiment. The loading configuration is shown in
Figure 4(a).

2.4. D OUBLE CANTILEVER BEAM LOADING CONFIGURATION

Our second loading configuration is a double cantilever beam type loading configuration used
for crack arrest experiments. In this configuration circular steel discs were attached to the
samples on either side of the seed crack (Figure 4). The samples were then loaded by dis-
placing the discs at a constant rate. Loading of this type is stable by energy considerations
(Freund, 1990) and can lead to crack arrest (Kalthoff et al., 1977). Displacement rates in
these experiments were 5 104 m/s. During the fracture process conditions can be assumed
to be quasi-static. Crack extension will thus unload the crack, meaning that the available
fracture energy is slowly decreased until crack extension is no longer possible, and the crack
arrests. Samples of PMMA, Homalite-100, and glass were loaded in this way. Crack arrest
was achieved in PMMA, and Homalite-100, but not in glass. The reason for this is most likely
that the tensile testing machine, with a compliance of 6.7 104 lb/in, is soft compared to
the glass sample, resulting in boundary conditions that are closer to constant force than to
constant displacement rate leading to unstable fracture.

3. Results

3.1. D ISSIPATION PROCESSES IN H OMALITE -100

It is well-known that Homalite-100 as well as many other brittle materials exhibits a sharp
rise in fracture energy once the crack velocity surpasses a certain threshold. As a result of this
rise in fracture energy cracks in most brittle materials never reach the theoretically predicted
sound speed. Since energy is not used to increase crack velocity it must be dissipated in other
ways. This dissipation process should leave behind traces that make it possible to account
for the energy introduced into the system. In PMMA it was shown by Sharon et al. that the
large increase in fracture energy could be explained by a crack tip instability that sets in at
a critical velocity (Sharon et al., 1995). Above the critical velocity the crack attempted to
branch off to the side with the side branches subsequently dying out, resulting in a series of
micro-cracks. The combined surface area of these micro-cracks in addition to the area of the
main crack accounts for nearly all the energy dissipated (Sharon et al., 1996). Here we want
to determine whether such micro-cracking, which is also well-known to exist in Homalite-
100 (Irwin et al., 1979; Kobayashi and Mall, 1978; Kobayashi and Dally, 1977; Dally, 1979;
142 J.A. Hauch and Marder

Figure 5. Velocity profile for three thin strip experiments in Homalite-100. At the top is a detailed view of the
crack initiation. The average acceleration during this stage is 2.7 108 m/s2 .

Figure 6. (a) Average Velocity vs. KI for all thin strip experiments in Homalite-100. (b) Fracture energy G and
normalized fracture energy G/Go vs. Average Velocity for the thin strip experiments in Homalite-100. The dotted
lines indicate the values reported by Dally for (1) minimum fracture energy, (2) onset of micro-cracking (3)
successful branching.

Ravi-Chandar and Knauss, 1984), can provide the same explanation for the increase in dis-
sipation in Homalite-100. Our findings show that this is not entirely the case, as there are
large increases in dissipation even before the onset of micro-cracking. We believe that this
dissipation occurs below the fracture surface where the material structure is modified by the
crack without leaving traces on the crack surface. We find some evidence for this; however,
thickness of the subsurface damage zone does not increase monotonically with energy flux as
we would expect.
The first step in the investigation was to determine the form of velocity as a function of
the stress intensity factor KI . For this purpose a series of samples was loaded in the strip
configuration. The values of KI and the fracture energy G were then obtained from (7), and the
average velocity was calculated using only the parts of the velocity records where the linear
Energy balance in dynamic fracture 143

resistance approximation in (2) holds. Typical velocity records for three different fracture
energies are shown in Figure 5. The results of all thin strip experiments in Homalite-100 are
summarized in Figure 6(a) and 6(b). Figure 6(a) shows the crack velocity as a function of stress
intensity factor. Our experimental results are compared to data obtained in a large series of
experiments done in the late 1970s (Irwin et al., 1979; Kobayashi and Mall, 1978; Kobayashi
and Dally, 1977). In a summary of these experiments by Dally (1979) effects due to specimen
geometry were averaged to produce the solid curve shown in Figure 6(a) and our data are
within the limits of this geometrical variation. Thus Figure 6(a) demonstrates the validity of
(7) for the thin strip geometry. Also indicated in the figure are the minimum stress intensity
factor for which propagation is possible, the stress intensity factor where attempted crack
branching is first observed, and the value of KI where successful crack branching occurred in
the specimen geometries used. An interesting feature of the thin strip geometry is that due to
the equilibrium with the side boundaries the instability is suppressed by reflected stress waves
and catastrophic branching prevented. This allows an extension of the KI vs.velocity curve to
much higher values of the stress intensity factor, in our case up to 1.78 MPa m.
Figure 6(b) recasts the same data in terms of fracture energy as a function of velocity. Again
the same three regions that were indicated in Figure 6(a) are indicated. The plot clearly shows
there are two transition regions. The first transition occurs approximately at G = 160 J/m2
and another one occurs at G = 500 J/m2 . For fracture energies between 80 J/m2 and 160J/m2
the fracture energy increases linearly with velocity at a rate of approximately 0.3 J s/m3 ; i.e.
small changes in fracture energy produce large changes in crack velocity. In the range from
160 J/m2 to 500 J/m2 the fracture energy still increases linearly with velocity but now at a
rate of 4 J s/m3 , a factor of 13 larger, but changes in G still produce significant changes in
velocity. Beyond energies of 500 J/m2 the slope is infinite and even large changes in energy
flux have no effect on the crack velocity up to fracture energies of about 1000 J/m2 at which
point increases in fracture energy affect the crack velocity again. The second transition point at
500 J/m2 coincides with the previously reported onset of the micro cracks. Beyond that point
all extra energy deposited in the system is dissipated by increasing the density and length of
the micro-cracks.
Figure 7 shows a series of transmission microscopy pictures of the crack at several different
values of the fracture energy. The pictures have a depth of field ranging from 20 m to 100 m
and thus show only a slice through the sample. Any particular micro-crack extends only a
fraction of the specimen thickness through the sample, such that the distribution of the micro-
cracks is highly three dimensional, with different layers of cracks emerging as the microscope
is focused at different depths. The images obtained by this method show no micro-cracks for
energies less than 500 J/m2 in accordance with the findings by Dally et al. By this point the
energy used to create new surface has increased by 600 percent over G0 , the minimum energy
at which crack propagation is possible. Observation of the fracture surface by eye reveals
that between the first and the second transition the fracture surface changes uniformly from a
mirror surface to a surface of uniform roughness. Detailed measurements of surface roughness
in PMMA however have shown that the extra surface created due to this mirror-mist transition
cannot account for the large increases in energy observed. Ravi-Chandar and Knauss explain
the extra dissipation in terms of parabolic surface markings. From our observations at high
fracture energies these surface markings are the surface signature of micro-cracks.
The first evidence of these parabolic markings was found at KI = 0.93 MPa m at an
energy 300 percent above the minimum, indicating that micro-cracking does exist for stress
intensity factors below the previously reported values. These micro-cracks are not observable
144 J.A. Hauch and Marder


energy (a) KI = 0.93 MPa m, G = 360 J/m
Figure 7. Development 2
of micro-cracking2 with increasing fracture
(b) KI = 1.11 MPa m, G = 513 J/m (c) KI = 1.23 MPa m, G = 630 J/m2 (d) KI = 1.33 MPa m,
G = 737 J/m2 (e) KI = 1.72 MPa m, G = 1233 J/m2 . The pictures show a top view of the crack which
propagated from left to right. To accommodate the rapidly increasing length of the micro-cracks the magnification
is decreased. There is no indication of micro-cracking in (a) from this view.

by transmission microscopy for two resons. First their total length is only on the order of 15
25 m (Figure 8(c)) with an extension into the sample of only a fracture of the length, which
makes them invisible due to diffraction of the light around the edge of the main crack, and due
to warping of the fracture surface which is larger than the extension of the micro-cracks into
the sample. Second their opening is very small, such that the total amount of light reflected by
them is negligible, making them
invisible. Nevertheless the density of these parabolic surface
markings at KI = 0.93 MPa m is quite low, making it unlikely that these are responsible for a
factor of 4.5 increase in dissipation. This indicates that other subsurface mechanisms besides
micro-cracking must be at play. A mechanism that appears to be important at these lower
energies is subsurface flaw growth as was proposed by Knauss and Ravi-Chandar (1984). A
large increase
in subsurface flaw size is noticeable as the stress intensity factor increases from
0.62 MPa m to 0.79 MPa m (Figures 8(a,b)). Using confocal microscopy it was determined
that these flaws exist only within a range of approximately 20 m of the surface, indicating
that the fracture process is responsible for their creation. The growth of these flaws also con-
stitutes a creation of new surface which will dissipate significant amounts of energy. As the
stress intensity factor increases further the density of these flaws decreases again as the surface
starts to break up due to river formations, in the vicinity of which the density of subsurface
flaws is noticeably less.
Energy balance in dynamic fracture 145

Figure 8. Development
of the fracture surface with increasing stress intensity factor, fracture
energy (a)
KI = 0.62 MPa m, G = 160 J/m (b) KI2= 0.79 MPa m, G =
2 263 J/m2 (c) KI = 0.93 MPa m, G = 360 J/m
2
(d) KI = 1.11 MPa m, G = 513 J/m (e) KI = 1.23 MPa m, G = 630 J/m2 (f) KI = 1.72 MPa m,
G = 1233 J/m2 . The parabolic markings become first visible in (c) and increase in size and density as the fracture
energy increases. They appear to be the surface signature of the micro-cracks in Homalite-100.

Figure 9. The figure shows the areas of the samples that are displayed in Figures 10 and 11. The detail on the
left is the sample edge that is shown in Figure 10, and the detail on the right is the sample corner that is shown in
Figure 11. The three important surfaces are labeled: (A) fracture surface, (B) original surface of the sample and
(C) surface created by subsequent fracture to reveal the structure inside.
146 J.A. Hauch and Marder

Figure 10. Edge viewof the fracture samples taken with SEM for increasing stress intensity factor, fracture energy

(a) KI = 0.51 MPa m, G = 108 J/m2 (b) KI = 0.62 MPa m, G = 160 J/m2 (c) KI = 0.79 MPa m,
G = 263 J/m2(d) KI = 0.93 MPa m, G = 360 J/m2 (e) KI = 1.23 MPa m, G = 630 J/m2 (f)
KI = 1.72 MPa m, GI = 1233 J/m2 . Figure 9 shows a sketch of the part of the sample that is imaged. In all
pictures the lighter surface on the top is the fracture surface (marked A in (a)) and the dark surface on the bottom
is the top surface (marked B in (a)) of the sample. All pictures indicate some form of subsurface activity; however
no clear trend is visible.


Figure 11. Three surface
SEM view for two values of the fracture energy (a) KI = 0.51 MPa m, G = 108 J/m2
(b) KI = 0.79 MPa m, G = 263 J/m2 . Figure 9 shows the part of the sample that is imaged. The surfaces are
marked in the following way: (A) fracture surface, (B) original surface of the sample and (C) surface created by
second fracture to reveal structure inside.

To investigate the subsurface structure further a series of Scanning Electron Microscope


pictures were taken of the edge of the fracture samples. The pictures shown in Figure 10
reveal that there is distinguishable structure below the fracture surface at all fracture energies.
The typical thickness of this damage zone is between 525 m. To determine if the observed
process zone exists everywhere underneath the fracture surface the samples were carefully
fractured one more time, this time perpendicular to the original fracture surface, to reveal
the structure below. This approach is intrinsically intrusive; nevertheless in several cases it
revealed structure underneath the fracture surface throughout the whole sample thickness.
Figure 11 shows SEM pictures of the edge where the three surfaces meet. Underneath the
fracture surface a filament structure exists that has a well defined boundary.
Energy balance in dynamic fracture 147

Creation of this damage zone should use up considerable amounts of fracture energy, but
it is unclear how one could quantify its contribution to dissipation. A first attempt would be
to measure the thickness of the damage zone as a function of fracture energy. However from
Figure 10 it is obvious that the thickness is not monotonically increasing with fracture energy
as one would expect. Furthermore the appearance of this damage zone changes significantly
with fracture energy. Thus dissipation in the damage zone may not be directly proportional
to the zone thickness. The pictures lead to the conclusion that the main source of dissipation
at all fracture energies is below the surface. However, we were puzzled to find that the width
of this damage zone does not appear to increase monotonically with energy until the onset of
micro-cracking and we cannot account for this observation.

3.2. T HE VELOCITY GAP

In the final part of this paper we want to turn our attention to the forbidden band of velocities
that is predicted by the analytical solution of the lattice models (Marder, 1993; Marder and
Gross, 1995). The prediction is that there is a forbidden band of velocities ranging from 0 m/s
up to about 20 percent of the Rayleigh wave speed. The only way for a crack velocity to pass
through this range is in the course of large acceleration or deceleration. Previous experiments
in PMMA in the thin strip configuration have shown that fracture initiation is always accom-
panied by a rapid acceleration stage up to about 20 percent of the Rayleigh wave speed. The
results shown in Figure 5 similarly show an initial rapid acceleration stage for Homalite-100,
in most cases all the way up to the steady state velocity. The experiments provide sufficient
time resolution to investigate crack initiation. The data in both materials clearly show that
there is an acceleration in excess of 2 108 m/s2 visible in the surface velocity at initation.
It is difficult however to attribute this behavior to any fundamental property since it may be
due to other fast processes that occur at the moment of initation. One possible explanation is
the three dimensionality of the plate, causing fracture to initiate preferentially in the middle
of the plate and subsequently growing outwards causing a sharp jump in the surface velocity
of the crack. Surface markings that grow spherically outwards at the beginning of the crack
support this picture in Homalite-100. Another possible explanation is crack tip blunting due
to plastic or visco-elastic flow in the vicinity of the crack tip. If fracture initiates from a blunt
seed-crack there must be a fast transition from a blunt crack to a sharp traveling crack. This
sharpening of the crack could also lead to a large acceleration visible in the surface velocity.
A way to alleviate both of these difficulties is to measure instead the arrest dynamics of
a crack, since a moving crack has a sharp crack front with an established three dimensional
profile. The crack arrest experiments were implemented in a Double Cantilever Beam config-
uration. In this configuration the unloading of the crack is slow compared to the time scales
available to the crack dynamics, allowing the crack always to seek out a steady state. In the
velocity gap picture presented by the molecular models a crack in such a configuration should
decelerate smoothly until some velocity on the order of 20 percent of the Rayleigh wave
speed and then rapidly arrest. Figures 12(a) and 12(b) show the results of such experiments
in PMMA and Homalite-100 respectively. Both experiments clearly show the characteristic
velocity jump at initiation that was also observed in the thin strip experiments. For the de-
celeration stage that begins almost immediately after the initial velocity jump the molecular
models predict some hysteresis, i.e. a slow deceleration down to a certain velocity lower than
the initial velocity jump, followed by a very rapid deceleration down to zero. In both, PMMA
and Homalite-100, the crack initally decelerates slowly with the deceleration growing at an
148 J.A. Hauch and Marder

Figure 12. Velocity records for experiments in the Double Cantilever Beam configuration. In PMMA and Homa-
lite-100 crack arrest was achieved. In both materials the crack appears to decelerate smoothly but with increasing
deceleration until it arrests. In glass crack arrest was not achieved in this configuration. However there is no sign
of an initial velocity jump. This can be attributed to the sharp seed cracks that can be generated in glass.

increasing rate as the velocity approaches zero. Nevertheless there is no clear transition point
between slow and rapid deceleration as would be expected from theory. The data clearly shows
that a forbidden band of velocities does not exist in PMMA and Homalite-100. Figure 12(c)
shows the result of an experiment in the same loading configuration in soda-lime glass. Due
to its low thermal shock resistance it is relatively easy to introduce sharp seed cracks with
an established crack-front into glass. Therefore glass does not suffer from the same problems
as the polymers during initiation. However crack arrest experiments in glass are extremely
difficult. No crack was successfully arrested in the DCB loading in glass, owing to the com-
pliance of the tensile testing machine used. The Rayleigh wave speed in soda-lime glass is on
the order of 3000 m/s. The forbidden band of velocities in glass should extend from zero up to
about 600 m/s. The data in Figure 12(c) shows smooth acceleration up to a velocity of 60 m/s.
This means that the forbidden velocity band picture also does not fit for glass. Thus, these
experiments indicate that the theoretical picture obtained in brittle crystals at low temperature
does not apply to amorphous materials at room temperature.

4. Conclusion

In our investigation we used a variation of the potential drop method for measuring the length
and velocity of running cracks. Instrumentation and data reduction were discussed in detail.
The advantage of this technique is that it is capable of yielding a direct measurement of
crack velocity that is sustainable at very high acquisition rates. Synchronization of the data
acquisition with initiation from a blunt seed crack is achieved with relative ease. Moreover
records of the full fracture process, including initiation, run, and arrest are attained at temporal
and spatial resolution that yield useful information about the dynamics of running cracks.
Motivated by the good qualitative agreement between the fracture of PMMA and the theory
of fracture of crystals at low temperatures, i.e. both display a micro-cracking instability which
Energy balance in dynamic fracture 149

constitutes the major mode of dissipation, we have investigated the fracture of Homalite-100
to determine if this same picture is applicable. In previous work it was found that micro-
cracking in Homalite-100 sets in at fracture energies which were a factor of 7 higher than the
minimum fracture energy at which crack propagation is possible. Although we have found
parabolic markings that indicate the existence of micro-cracks at fracture energies signifi-
cantly lower than that, a factor of 4.5 above the minimum fracture energy, it is clear that
in Homalite-100 other modes of dissipation must be significant. Post mortem observations
on the fracture samples with a Scanning Electron Microscope show that at all energies there
exists a damage zone with a thickness between 1025 m underneath the fracture surface. The
material structure in this zone is modified by filamentation and void growth that at low fracture
energies remain hidden below a perfectly smooth fracture surface. At low fracture energies this
damage zone is probably the main source of the dissipation, while at higher energies a complex
three dimensional set of micro-cracks evolves and becomes the main source of dissipation.
We are puzzled by the observation that the thickness of the damage zone does not increase
monotonically with fracture energy, and it is unclear how it evolves until the onset of micro-
cracking. It appears that the fracture processes of Homalite-100 are much more complex and
difficult to describe than the processes in PMMA. We also investigated the existence of a
forbidden band of velocities that exists in the molecular models. In PMMA and Homalite-100
we observe a velocity jump during crack initition that we attributed to crack-tip blunting or
three dimensional effects. However our experiments showed that no such forbidden band of
velocities exists in amorphous materials at room temperature.

Acknowledgements

This project was supported by a TARP (Texas Advanced Research Program) grant issued
for the investigation of dynamic fracture. Special thanks to the Cell Research Institute of the
University of Texas at Austin for the use of their Scanning Electron Microscope, and John
Mendenhall, Barbara Goettgens for their time.

References

Abraham, F.F., Brodbeck, D., Rafey, R.A. and Rudge, W.E. (1994). Instability dynamics of fracture: A computer
simulation investigation. Physical Review Letters 73(2), 272275.
Bguelin, P., Stalder, B. and Kausch, H.H. (1983). Application of a new velocity gage to fracture at high velocities.
International Journal of Fracture 23, R7R10.
Buck, O. (1989). Recent advances in fracture mechanics testing. In Fracture Mechanics: Microstructure and
Micromechanisms. (Edited by S.V. Nair, J.K. Tien, R.C. Bates, and O. Buck) ASM International, Metals Park,
OH, 3186.
Cotterell, B. (1965). Velocity effects in fracture propagation. Applied Materials Research 4, 227232.
Cudr-Mauroux, N., Kausch, H.H., Cantwell, W.J. and Roulin-Moloney, A.C. (1991). High speed crack
propagation in bi-phase materials: An experimental study. International Journal of Fracture 50, 6777.
Dally, J.W. (1979). Dynamic photoelastic studies of fracture. Experimental Mechanics 19, 349361.
Doyle, M. A mechanism of crack branching in polymethylmethacrylate and the origin of bands on the surface of
fracture. Journal of Materials Science 18, 687702.
Dulaney, E.N. and Brace, W.F. (1960). Velocity behavior of a growing crack. Journal of Applied Physics 31,
22332266.
Field, J.E. (1971). Brittle fracture: its study and application. Contemporary Physics 12, 131.
Fineberg, J., Gross, S.P., Marder, M.P. and Swinney, H.L. (1991). Instability in dynamic fracture. Physical Review
Letters 67, 457460.
150 J.A. Hauch and Marder

Fineberg, J., Gross, S.P., Marder, M.P. and Swinney, H.L. (1992). Instability in the propagation of fact crack.
Physical Review B 45, 51465154.
Freund, L.B. (1990). Dynamic Fracture Mechanics. Cambridge University Press, New York.
Gilman J.J., Knudsen, C. and Walsh, W.P. (1958). Cleavage cracks and dislocations in LiF crystals. Journal of
Applied Physics 6, 601607.
Griffith, A.A. (1920). The phenomena of rupture and flow in solids. Mechanical Engineering A221, 163198.
Gross, S.P. Fineberg, J., Marder, M.P., McCormick, W.D. and Swinney, H.L. (1993). Acoustic emissions from
rapidly moving cracks. Physical Review Letters 71(19), 31623165.
Gumbsch, P. (1995). An atomistic study of brittle fracture: Toward explicit failure criteria from atomistic modeling.
Journal of Materials Research 10(11), 28972907.
Hall, E.O. (1953). The brittle fracture of metals. Journal Mechanics and Physics Solids 1, 227233.
Hull, D. and Beardmore, P. (1966). Velocity of propagation of cleavage cracks in tungsten. International Journal
of Fracture Mechanics 2, 468487.
Irwin, G.R., Dally, J.W., Kobayashi, T., Fourney, W.L., Etheridge, M.J. and Rossmanith, H.P. (1979). On the
determination of the a-k relationship for birefringent polymers. Experimental Mechanics 19(4), 121128.
Kalthoff, J.F., Beinert, J. and Winkler, S. (1977). Measurements of dynamic stress intensity factors for fast running
cracks in double- cantilever-beam specimens. In Fast Fracture and Crack Arrest (Edited by G.T. Hahn and
M.F. Kanninen) ASTM STP 627, 161176.
Kerkhof, F. (1973). Wave fractographic investigations of brittle fracture dynamics. In Dynamic Crack Propagation
(Edited by G.C. Sih) Noordhoff International Publishing, Leyden, 329.
Knauss, W.G. and Ravi-Chandar, K. (1985). Some basic problems in stress wave dominated fracture. International
Journal of Fracture 27, 127143.
Kobayashi, T. and Dally, J.W. (1977). Relation between crack velocity and the stress intensity factor in birefringent
polymers. In Fast Fracture and Crack Arrest (Edited by G.T. Hahn and M.F. Kanninen) ASTM STP 627, 718.
Kobayashi, A.S. and Mall, S. (1978). Dynamic fracture toughness of Homaltie-100. Experimental Mechanics
18(1), 1118.
Kobayashi, A., Ohtani, N. and Sato, T. (1974). Phenomenological aspects of viscoelastic crack propagation.
Journal of Applied Polymer Science 18, 16251638.
Liu, X. and Marder, M.P. (1991). The energy of a steady-state crack in a strip. Journal of Mechanics and Physics
Solids 39, 947961.
Marder, M.P. (1993). Simple models of rapid fracture. Physica D 66, 125134.
Marder, M.P. and Gross, S.P. (1995). Origin of crack tip instabilities. Journal of Mechanics and Physics of Solids
43, 148.
Marder, M.P. and Liu, X. (1993). Instability in Lattice fracture. Physical Review Letters 71, 24172420.
Mott, N.F. (1947). Brittle fracture in mild steel plates. Engineering 165, 1618.
Ravi-Chandar, K. and Knauss, W.G. (1982). Dynamic crack-tip stress under stress wave loadinga comparison of
theory and experiment. International Journal of Fracture 20, 209222.
Ravi-Chandar, K. and Knauss, W.G. (1984a). An experimental investigation into dynamic fracture: I. Crack
initiation and arrest. International Journal of Fracture 25, 247262.
Ravi-Chandar, K. and Knauss, W.G. (1984b). An experimental investigation into dynamic fracture: II. Microstruc-
tural aspects. International Journal of Fracture 26, 6580.
Ravi-Chandar, K. and Knauss, W.G. (1984c). An experimental investigation into dynamic fracture: III. On steady-
state crack propagation and crack branching. International Journal of Fracture 26, 141154.
Ravi-Chandar, K. and Knauss, W.G. (1984d). An experimental investigation into dynamic fracture: IV. On the
interaction of stress waves with propagating cracks. International Journal of Fracture 26, 189200.
Schardin, H., Mucke, L., Struth, W. and Rhein, W.A. (1955). Cracking velocity of glasses. The Glass Industry
36(3), 133138.
Sharon, E., Gross, S.P. and Fineberg, J. (1995). Local crack branching as a mechanism for instability in dynamic
fracture. Physical Review Letters 74, 51465154.
Sharon, E., Gross, S.P. and Fineberg, J. (1996). Energy dissipation in dynamic fracture. Physical Review Letters
76(12), 21172120.
Sinclair, J.E. and Lawn, B.R. (1972). An atomistic study of cracks in diamond-structure crystals. Proceedings of
the Royal Society A 329 83103.
Stalder, B., Bguelin, P. and Kausch, H.H. (1983). A simple velocity gauge for measuring crack growth.
International Journal of Fracture 22, R47R54.
Energy balance in dynamic fracture 151

Stalder, B., Bguelin, P., Roulin-Moloney, A.C. and Kausch, H.H. (1989). The graphite gauge and its application
to the measurement of crack velocity. Journal of Materials Science 24, 22622274.
Stalder, B. and Kausch, H.H. (1985). The use of a velocity gauge in impact testing of polymers. Journal of
Materials Science 20, 28732881.
Stroh, A.N. (1957). A theory of the fracture of metals. Philosophical Magazine 6, 418465.
Thomson, R., Hsieh, C. and Rana, V. (1971). Lattice trapping of fracture cracks. Journal of Applied Physics 42(8),
31543160.
Wang, X.M. and Kim, H.S. (1993). Continuous measurement of crack growth in plastics. Journal of Materials
Science Letters 12, 357358.
Weimer, R.J. and Rogers, H.C. (1977). A high-speed digital technique for precision measurement of crack ve-
locities. In Fast Fracture and Crack Arrest (Edited by G.T. Hahn and M.F. Kanninen) ASTM STP 627,
359371.
Yoffe, E.H. (1951). The moving griffith crack. Philosophical Magazine 42, 739750.
Zhou, S.J., Carlsson, A.E. and Thomson, R. (1994). Crack blunting effects on dislocation emission from cracks.
Physical Review Letters 72 852855.

You might also like