You are on page 1of 9

Engineering Fracture Mechanics 76 (2009) 1268–1276

Contents lists available at ScienceDirect

Engineering Fracture Mechanics


journal homepage: www.elsevier.com/locate/engfracmech

Determination of dynamic fracture parameters using a semi-circular


bend technique in split Hopkinson pressure bar testing
R. Chen a,b, K. Xia a,*, F. Dai a, F. Lu b, S.N. Luo c
a
Department of Civil Engineering and Lassonde Institute, University of Toronto, ON, Canada M5S 1A4
b
College of Science, National University of Defense Technology, Changsha, Hunan 410073, PR China
c
Physics Division, Los Alamos National Laboratory, Los Alamos, New Mexico 87545, USA

a r t i c l e i n f o a b s t r a c t

Article history: Fracture initiation toughness, fracture energy, fracture propagation toughness, and fracture
Received 4 November 2008 velocity are key dynamic fracture parameters. We propose a method to simultaneously
Received in revised form 4 February 2009 measure these parameters for mode-I fractures in split Hopkinson pressure bar (SHPB)
Accepted 5 February 2009
testing with a notched semi-circular bend (SCB) specimen. The initiation toughness is
Available online 13 February 2009
obtained from the peak load given dynamic force equilibrium. A laser gap gauge (LGG) is
developed to monitor the crack surface opening displacement (CSOD) of the specimen,
Keywords:
from which the fracture velocity and the fracture energy can be calculated. The feasibility
Dynamic fracture mechanics
Dynamic initiation toughness
of this methodology for coarse-grained solids is demonstrated with the SHPB-SCB experi-
Dynamic propagation toughness ments on Laurentian granite.
Fracture velocity Ó 2009 Elsevier Ltd. All rights reserved.

1. Introduction

Dynamic fracture plays a vital role in geophysical processes and engineering applications (e.g., earthquakes, airplane
crashes, projectile penetrations, rock bursts and blasts). Accurate measurements of dynamic fracture parameters are prere-
quisite for understanding mechanisms of fracture and also useful for engineering applications. For brittle materials such as
rocks, one can not simply use the standard methods of fracture measurement developed for metals. Special sample geom-
etries have been adopted for fracture toughness measurements of ceramics, rocks and concretes [1–3]. For example, Chong
and Kuruppu used the semi-circular bend (SCB) specimen for rocks [4]. This sample geometry is advantageous for convenient
sample preparation (directly from rock cores) and for relative ease of experimentation, and thus used, here for dynamic frac-
ture experiments.
Most experimental studies on fracture of rocks and other brittle materials are carried out under quasi-static loading con-
ditions. However, in many applications as mentioned above, materials are stressed dynamically. Limited attempts have been
made to measure the dynamic initiation fracture toughness of brittle solids with split Hopkinson pressure bar (SHPB). For
example, Zhang et al. conducted first measurements on rocks using short-rod specimens [5,6]; the inertial effect associated
with the stress wave loading as demonstrated by Böhme and Kalthoff [7] was ignored. One way to remedy this inadequacy is
to combine experimental measurements with numerical simulations [8], but this process is rather tedious. Recently, Owen
et al. observed that the stress intensity factors obtained by directly measuring the crack tip opening are consistent with those
calculated with the quasi-static equation when the dynamic stress equilibrium of the specimen is roughly achieved in split
Hopkinson tension bar testing [9]. This concept was subsequently adopted by Weerasooriya et al. for ceramics in SHPB
testing with four-point bend specimens [10].

* Corresponding author. Tel.: +1 416 978 5942; fax: +1 416 978 6813.
E-mail address: kaiwen.xia@utoronto.ca (K. Xia).

0013-7944/$ - see front matter Ó 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.engfracmech.2009.02.001
R. Chen et al. / Engineering Fracture Mechanics 76 (2009) 1268–1276 1269

It is highly desirable to develop methods for measuring dynamic fracture parameters in a single experiment, including
facture initiation toughness, fracture energy, fracture propagation toughness, and fracture velocity. Previous studies focus
on the initiation toughness. The fracture energy and the fracture propagation toughness are directly related to the energy
consumption during dynamic failures. To our knowledge, there is only one attempt to measure the dynamic propagation
toughness of rocks [11], where an array of strain gauges was used to measure the strain field associated with fracture prop-
agation. For dynamic fracture, the shrinkage of the domain of small scale yielding may lead to significant error for methods
based only on the singular term of a stress field (e.g., the strain gauge and caustics methods) [12]. Indeed, six terms of expan-
sion of the stress field is required to fit the photoelastic fringe patterns [13].
Fracture propagation toughness and fracture velocity can be better measured with optical methods for transparent poly-
mers or polished metals [9,13]. The concept of optical techniques in SHPB testing was initiated by Griffith and Martin [14],
who used white light to monitor the displacements at the end faces of a cylindrical specimen. Tang and Xu measured the
crack surface opening displacement (CSOD) using a line source of white light [15], and took the turning point of the CSOD
history as the fracture initiation time. Zhang et al. used the Moiré method to monitor the CSOD of short-rod specimens, and
assumed that the peak point of the opening velocity curve obtained from CSOD corresponds to the onset of fracture [5].
Nonetheless, these assumptions were not sufficiently justified. Ramesh and Kelkar adopted a line laser source to measure
the velocity history of flyer in planer impact [16]. Later, Ramesh and Narasimhan used this technique to measure the radial
expansion of specimens in SHPB tests [17]. In this study, a laser gap gauge (LGG) is used to monitor CSOD of a SCB specimen
during SHPB testing.
The residue kinetic energy in the fragments can be obtained from the fragment velocity, which is the temporal derivative
of the CSOD history. Given the residual fragment kinetic energy and total energy consumption (deduced from the strain
gauge signals), both the initiation fracture toughness and the propagation fracture toughness are determined. A similar
method was attempted by Zhang et al. who used a high-speed camera to estimate the fragment velocities [6]. The fracture
toughness was not explicitly presented in their work and no information on the fracture velocity could be obtained using
their technique. We show that using the CSOD data measured with LGG, we can also determine the completion time of
the fracture. Because we have achieved dynamic force balance in all of our tests, the peak in the loading history corresponds
to the onset time of the fracture. The total fracture growth length is measured from the sample geometry. We can thus cal-
culate the averaged fracture velocity for the SCB rock specimen.
In this work, we report the methodology of measuring all the four fracture parameters in a single SHPB test with a SCB
specimen, and demonstrate its application to a granitic rock. The overall experimental setup, including the SCB geometry,
sample preparation, LGG and SHPB testing, are discussed in Section 2. The techniques and principles for measuring the frac-
ture initiation toughness, energy, propagation toughness and velocity are detailed in Section 3. Section 4 presents an illus-
trative example of its application. Main conclusions are summarized in Section 5.

2. Experimental setup for dynamic fracture of a SCB specimen under SHPB loading

Dynamic fracture tests on rock materials usually resort to compression-induced tension in order to avoid pre-mature fail-
ure due to gripping in purely tensile testing. Chong and Kuruppu [4] adopted a SCB rock specimen with a single edge notch to
measure fracture toughness in a compression setting. The geometry of a SCB specimen is shown in Fig. 1: its radius is R and
thickness is B, the depth of the notch is a, and the span of the supporting pins is S. The force applied on the side without any
supporting pins is P1; the force is P2 on the other side with two pins, and P2/2 is exerted on each pin. Sufficient crack tip
sharpness is necessary for accurately measuring fracture initiation toughness [18–20]. For an ideal crystal, the naturally
formed crack has a finite thickness of the order of atomic spacing; for a polycrystalline solid, the thickness is comparable
to its grain size. The thickness of an intergranular pre-crack formed by fatigue or other methods is on the order of the char-
acteristic material length (e.g., the average grain size in a polycrystalline solid). As discussed by Lim et al. and references

Striker Incident bar Transmitted bar

P1 a P2
S

Cylindrical lens Crack gauge Sample Collecting lens

Gap
Laser B Detector

Fig. 1. Schematics of the split Hopkinson pressure bar (SHPB) system, the semi-circular bend (SCB) sample, and laser gap gauge (LGG) system.
1270 R. Chen et al. / Engineering Fracture Mechanics 76 (2009) 1268–1276

therein [21], pre-cracking may not be necessary for certain rocks if the notch is sufficiently small (<0.8 mm). In our exper-
iments, we first make a 1 mm wide notch in the semi-circular rock disc and then sharpen the tip with a diamond wire saw to
achieve a tip diameter of 0.5 mm. The average grain size is about 0.5 mm [22], so the diameter of the crack tip is similar to
the thickness of naturally formed cracks.
The compression in our dynamic fracture tests is achieved with a 25 mm diameter SHPB system (Fig. 1). The bars are
made from Maraging steel, with a high yield strength of 2.5 GPa. The length of the striker bar is 200 mm. The incident (input)
bar is 1500 mm long and the strain gauge station is 787 mm from the impact end of the bar. The transmitted (output) bar is
1000 mm long and the stain gauge station is 522 mm away from the specimen. The striker bar is launched by a low pressure
gas gun to impact on the incident bar, resulting in an incident stress wave. The incident pulse propagates along the incident
bar before it hits the sample, resulting in the reflected and transmitted stress waves. An eight-channel Sigma digital oscil-
loscope by Nicolet is used to record and store the strain signals collected from a home-made Wheatstone bridge circuits after
amplification. The sampling rate used is 10 MHz and at this rate, the resolution of the scope is 12-bit. Denote the incident
wave, reflected wave and transmitted wave with subscripts i, r and t, respectively. The loading forces on the two ends of the
specimen induced by the SHPB are [23]:
P1 ¼ AEðei þ er Þ; P2 ¼ AEet ð1Þ
where E is Young’s modulus of the bar material, A is the cross-sectional area of the bar and e denotes strain.
The LGG system is composed of a diode line laser, a laser holder, a cylindrical lens, a collecting lens and a light detector
(model PDA100A from Thorlabs) (Fig. 1). The laser operates at 670 nm with a 5 mW output power. It has a large field depth
and minimal variations in thickness across the line length. The line is 30 lm thick at 185 mm away, and the divergence angle
is 5°. A cylindrical lens is used to achieve a parallel laser sheet. The plano-convex cylindrical lens is made from coated BK7
glass. The high performance multilayer anti-reflection coatings have an average reflectance of less than 0.5% (per surface) in
the wavelength range of 650–1050 nm. The light detection part consists of a collecting lens and a photodiode light detector.
The collecting lens focuses the incoming laser light into the photodiode detector, which is placed near its focal point. A nar-
row-band-pass filter centered at 670 nm is placed in front of the detector window. The photodiode detector output is pre-
amplified, and the optoelectronics and the preamplifier together have a bandwidth of 1.5 MHz. The output voltage of the
detector is proportional to the total amount of laser light collected. The whole system has a noise level less than 0.4 mV.
The LGG system monitors the change of the width of the laser sheet that passes through a gap. In SHPB testing on an SCB
specimen, LGG is mounted perpendicular to the bar axis and the laser passes through the notch in the center of the specimen
(Fig. 1). As the notch opens up, the amount of light passing through increases, leading to higher voltage output from the
detector. The voltage is linearly proportional to the gap width and thus CSOD.

3. Measurement techniques and principles

3.1. Calibration of the LGG

Calibration of LGG is conducted under both static and dynamic conditions. For static calibration, we use a set of high pre-
cision gauges to partly block the probe laser (Fig. 2 inset). The blocking width ranges from 0 to 10 mm at a step of 0.1 mm. A

Measurement
3 Linear fitting
Δd (mm)

2
X

1
d
Collimated beam Gap gauge
0
0.0 0.2 0.4 0.6 0.8 1.0
ΔU (V)
Fig. 2. Static calibration of the LGG system: schematic setup (inset) and the calibration result.
R. Chen et al. / Engineering Fracture Mechanics 76 (2009) 1268–1276 1271

specific blocking width (d) corresponds to a light-passing width Dd and a certain amount of voltage reading (DU) in the
detector output (Fig. 2). The Dd  DU curve exhibits a good linearity, indicating the high uniformity of the laser sheet:
Dd ¼ kDU ð2Þ
where k = 4.08 mm/V is the calibration parameter of the LGG system. Denote the standard deviation as r, and the error prop-
agation follows from Eq. (2) as:
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
rDd =Dd  ðrk =kÞ2 þ ðrDU =DUÞ2 ð3Þ

Here rk = 0.03 mm/V, and rDU = 0.4 mV. As an example, for DU = 0.25 V, Dd = 1.02 mm and the relative error of the dis-
placement measurement is around 0.8%.
The dynamic calibration is carried out with a single Hopkinson bar. In order to get higher cutoff frequency, a miniaturized
6.35 mm diameter Hopkinson bar is utilized to calibrate the LGG system. The cutoff frequency fc of a long bar is [24]:
pffiffiffi
2C
fc ¼ 0:3 ð4Þ
2pR0 m
where R0 is the bar radius, m and C are Poisson’s Ratio and elastic wave velocity, respectively. fc is 400 kHz for the 6.35 mm
diameter bar, much higher than 100 kHz for the 25 mm diameter bar.
We impact one end of the incident bar with the striker bar and use LGG to monitor the motion of the other end. When the
compressive wave arrives at the free end of the incident bar, it is reflected as a tensile wave. The incident and reflected waves
measured from the strain gauge glued on the incident bar. The strain signals are corrected for the travel time from the gauge
to the free end of bar. Then the displacement of the free end follows as [23]:
Z t
Dd ¼ ðei þ er ÞCds ð5Þ
0

where t denotes time. With the LGG output, we also calculate the displacement of the free end of the incident bar using Eq.
(2) for comparing the values obtained with Eq. (5). The comparison shows excellent agreement (Fig. 3), demonstrating that
the LGG system has a wide bandwidth sufficient for valid measurements in SHPB.

3.2. Dynamic force balance

The pulse shaper technique is employed to achieve dynamic force balance in the specimen during the experiment, i.e.,
P1 = P2. This technique was recently discussed in detail by Frew et al. for SHPB compressive tests of brittle materials [25].
In a traditional SHPB test, the incident wave with a sharp rising edge may initiate undesired damage to the sample upon
impact. Consequently, the forces on both sides of the specimen are not be the same, likely resulting in misinterpretation
of data [10]. We use a C11000 copper disc to shape the incident wave from the rectangular shape to a ramped wave. In addi-
tion, a rubber disc is placed in front of the copper shaper to reduce the rising slope of the incident pulse. This combined pulse
shaping technique was also used by other researchers [26,27].

1.0

0.8 Using Eq.2


Using Eq.5
Displacement (mm)

0.6

0.4

Incident bar
0.2
X
Laser beam
0.0
0 100 200 300 400
t (μs)
Fig. 3. Dynamic calibration of the LGG system: schematic setup (inset), and a typical dynamic test result compared to the predictions by Eq. (2).
1272 R. Chen et al. / Engineering Fracture Mechanics 76 (2009) 1268–1276

75

50

25

Force (kN)
0

-25 In
Tr
-50 Re
In+Re
-75
0 50 100 150 200 250
t (μs)
Fig. 4. Dynamic force balance during a typical SHPB test with the SCB specimen.

Fig. 4 shows the forces on both ends of the specimen in a typical test. From Eq. (1), the dynamic force on one side of the
specimen P1 is proportional to the sum of the incident (In) and reflected (Re) stress waves, and the dynamic force on the
other side P2 is proportional to the transmitted (Tr) stress wave. For SCB tests, P2 is equally shared by two pins. It can be seen
from Fig. 4 that the dynamic forces on both sides of the specimens are almost identical during the whole dynamic loading
period. The inertial effects are thus eliminated because there is no global force difference in the specimen to induce inertial
force. Consequently, the inertial effects are negligible in such cases and we can then perform quasi-static analysis [10] (see
below).

3.3. Loading history and fracture velocity

Fig. 5 shows a typical loading history (P2) and the corresponding CSOD history during a SHPB-SCB test. Because the dy-
namic force balance is achieved, the peak point of the loading (A) corresponds to the fracture initiation in the specimen, as in
a quasi-static experiment. The temporal derivative of the CSOD history is the crack surface opening velocity (CSOV) history.
CSOV increases with time and then approaches a terminal velocity of v = 13.9 m/s at the turning point B. The two vertical
lines passing through points A and B divide the whole deformation period into three stages I–III. We believe that in stage
I the crack opens up elastically, in stage II the crack propagates dynamically, and in stage III the fracture separates the sample
into two pieces and the two fragments rotate away from each other. The separation velocity of the two fragments (normal to

20 5
CSOD (mm) /CSOV (10 m/s)

8μs
16 4
3μs
Crack gauge
Force (kN)

12 3
20μs 22μs
Force
8 CSOD A 53μs 2
CSOV
B

4 1
13.9m/s

0 0
0 50 100 150 200 250 300
t (μs)
Fig. 5. Typical loading history and CSOD history of the SCB specimen tested in SHPB. Inset: the crack gauge signals at three locations.
R. Chen et al. / Engineering Fracture Mechanics 76 (2009) 1268–1276 1273

the bar axis) is approximately the terminal velocity in CSOV (for small angle of rotation in stage III), and doubles the frag-
ment velocity.
The crack propagation process lasts about DtAB = 53 ls as seen from CSOD and CSOV. Given the crack distance
Ls = R  a = 16 mm for this test, we estimate the average crack growth velocity vf to be about 300 m/s. We also use crack
gauges to estimate the fracture propagation velocity. Three cracks are mounted on the specimen (Fig. 1), separated by
Dl1 = 5.36 mm and Dl2 = 7.81 mm. The time separations between the arrivals of the fracture signals are Dt1 = 20 ls, and
Dt2 = 22 ls, respectively (Fig. 5 inset). Thus the corresponding fracture velocities are v1 = 268 m/s, and v2 = 355 m/s. The frac-
ture velocity appears to increase as the crack propagates during dynamic loading. The first gauge is cemented at about 2 mm
away from the crack tip in order to avoid interfering the crack initiation. So there is 8 ls delay between the crack initiation
and the breaking of the first crack gauge. The fracture velocity as measured with LGG is consistent with the crack gauge
results. One advantage of the LGG is that it is a non-contact method.

3.4. Initiation fracture toughness and loading rate

Based on the ASTM standard E399-06e2 for rectangular three-point bending sample [28], we propose a similar equation
for calculating the stress intensity factor for mode-I fracture in current SCB specimen:
PðtÞS a
K I ðtÞ ¼ 3=2
Y ð6Þ
BR R
Here P(t) is the time-varying loading force. Y(a/R), a dimensionless geometry factor, is a function of the crack geometry,
which can be calculated with a standard finite element software package (e.g., ANSYS).
As mentioned earlier, the forces applied on both sides of the sample during our SHPB tests are identical, i.e, F = P1/2 =
P2/2 = P/2 (Fig. 6). The inertial effects are eliminated because there is no global force difference in the specimen to induce
inertial forces [10]. We conduct finite element analysis to relate the far-field loading to the local stress intensity factor at
the crack tip for a given specimen geometry. This process is called numerical calibration. Taking advantage of the symmetry
of the problem, half-crack model is used to construct the finite element model. PLANE82 (eight-node) element is used in the
analysis. To better simulate the stress singularity of r1/2 near the crack tip (r is the radius to the crack tip), 1/4 nodal element
(singular element) [29] is applied to the vicinity of the crack tip in meshing the finite element model (Fig. 6). The Young’s
modulus is 92 GPa and the Poisson’s ratio is 0.21 for Laurentian granite used in our test [30]. The load is set as the boundary
stresses at the left and right edge of the model plate while the lower edge of the model has the symmetric boundary con-
dition. The resulting loading at the main crack is mode I. For a given load P, KI can be obtained from the finite element anal-
ysis. And the geometry factor Y(a/R) for a given sample geometry follows from Eq. (6) as:
a K I BR3=2
Y ¼ ð7Þ
R PS
Since Y(a/R) is calibrated, K I ðtÞ is directly calculated from Eq. (6) for loading history P(t). The initiation fracture toughness K dIC
corresponds to the peak point of the loading Pmax. There is a approximately linear region in KI(t), and its slope is taken as the
fracture loading rate, with which K dIC varies.

3.5. Propagation fracture energy and toughness

A high-speed camera (Photron Fastcam SA1) is used to monitor the fracture initiation and propagation process as well as
the trajectories of the fragments. The high-speed camera is placed perpendicular to the SHPB and specimen. Images are re-
corded at an interframe interval of 8 ls; the sequence shown in Fig. 7 represents only the frames of representative features.
The first two images show the pre-fabricated notch and the crack opening can be barely seen. The opening of the SCB crack

ω F

v/2
Δl
F
A 1 3
4 4

Symmetry LGG Crack tip

Fig. 6. Schematic of the configuration for finite element analysis, and the quarter point element on the crack tip.
1274 R. Chen et al. / Engineering Fracture Mechanics 76 (2009) 1268–1276

Fig. 7. Selected high-speed camera images showing the fracture and fragmentation of a SCB specimen.

becomes visible at t > 40 ls. At 80 ls, the SCB specimen is split completely into two fragments. The fragments then rotate
about the contact point between the specimen and the incident bar (point A in Fig. 6). The rotation angle of the fragment
is measured to be 9° at 160 ls, 21° at 480 ls, and 32° at 800 ls. This indicates that the angular velocity of the fragments
is almost constant during the period (about 314 rad/s), and the motion of the fragments is rotational.
The high-speed camera imaging indicates that the fragments rotate around the axis along the loading point A (Fig. 6). The
LGG system measures CSOD and the fragment angular velocity can be deduced. The linear velocity of the two rotating frag-
ments at the LGG point is approximately the terminal velocity in the CSOV curve (Fig. 5). The distance between the LGG and
the rotating axis Dl = 18 mm, so the angular velocity x = v/2/Dl = 386 rad/s for the shot shown in Fig. 6, in excellent agree-
ment with the result obtained from high speed imaging.
We next use the energy conservation principle to calculate the propagation fracture energy and fracture toughness. A
similar method was used by Zhang et al. who used a high-speed camera to estimated the fragment residual velocities [6].
The elastic energy carried by a stress wave is [31]:
Z t
W¼ Ee2 ACds ð8Þ
0

The total energy absorbed by the specimen then is DW = Wi  Wr  Wt. Part of the total energy absorbed is used to create
new crack surfaces, called the total fracture energy (WG); the other part remaining in the fragments as the residue kinetic
energy (K). That is, DW = WG + K. For the rotating fragments (Fig. 6), the moment of inertia is I, and the total rotational kinetic
energy is K = Ix2/2, where the fragment angular velocity x is estimated from the CSOD data. The average propagation frac-
ture energy is Gc = WG/Ac, where Ac is the area of the crack surfaces created. The average dynamic propagation fracture tough-
ness is:
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
K dP
IC ¼ Gc E=ð1  m2 Þ ð9Þ

where E and m are the Young’s modulus and Poisson’s ratio of the specimen respectively. Here we assume the plain strain
condition.

4. Application to Laurentian granite

To demonstrate the feasibility of the above methodology for measuring the dynamic fracture parameters, we have per-
formed full SHPB tests on Laurentian granite. The material is chosen because its mineralogical and mechanical character-
istics are well known [22]. Laurentian granite is from the Laurentian region of Grenville province of the Precambrian
Canadian Shield, north of St. Lawrence and north-west of Quebec City, Canada. The dominant constituents are feldspar
(60%) and quartz (33%). The mineral grain size of Laurentian granite varies from 0.2 to 2 mm with average grain sizes
of 0.5 mm and 0.4 mm for quartz and feldspar, respectively. Biotite grain size is of the order of 0.3 mm and constitutes
3–5% of this rock. Since Laurentian granite is fine-grained, we can use relatively small specimens for testing its bulk
fracture properties.
R. Chen et al. / Engineering Fracture Mechanics 76 (2009) 1268–1276 1275

25
Notch depth ~ 4 mm
d
KIC
20
dP

Toughness (MPa m )
KIC

1/2
15

10

0
20 40 60 80 100
1/2 -1
KI (GPa m s )

Fig. 8. The effect of loading rate on the initiation and propagation fracture toughnesses.

20
KIC (MPa m )
1/2

16
dP

12

300 450 600 750 900


vf (m/s)

Fig. 9. The variation of propagation fracture toughness with fracture velocity.

Rock cores with a nominal diameter of 40 mm are drilled from Laurentian granite blocks, and then sliced into disks with
an average thickness of 16 mm. A circular diamond saw is used to cut a disk into two half-disks and to make a 4 mm by 1 mm
notch on each half-disk. After the SHPB testing, the recovered sample is clearly split into two pieces. For these samples,
a = 4 mm, R = 40 mm, Y(a/R) = 0.086, and S = 20.1 mm. These values along with Pmax are substituted into Eq. (6) to calculate
K dIC . Other parameters are also determined with the methodology discussed in the previous section.
Fig. 8 shows the measured initiation and propagation fracture toughnesses at different loading rates; both of them in-
crease linearly with increasing loading rates. The propagation fracture toughness also increases with the fracture velocity
(Fig. 9). At the highest fracture velocity (850 m/s), the fracture toughness value is 18.5 MPa m1/2, about twice of those at
slower fracture velocities near 300 m/s.

5. Conclusion

We propose a SHPB testing technique with a notched semi-circular bend specimen to measure the dynamic fracture
parameters of mode-I fracture, including the initiation fracture toughness, average fracture energy, average propagation
toughness, and average fracture velocity of brittle materials. The initiation fracture toughness is obtained from the peak
loading force and the static analysis provided that the force balance is achieved during the SHPB testing. We develop a laser
gap gauge system to measure the crack surface opening displacement history. From this and SHPB measurements, the
1276 R. Chen et al. / Engineering Fracture Mechanics 76 (2009) 1268–1276

fracture energy, propagation fracture toughness and fracture velocity can be determined. This methodology is demonstrated
with experiments on Laurentian granite. Both the initiation and propagation toughnesses of this brittle solid are loading rate
dependent. The propagation toughness is larger than the initiation toughness for a given test. The propagation fracture
toughness is shown to increase with the fracture velocity. This technique is readily implementable and can be applied to
investigating fracture mechanics of brittle materials with coarse grains.

Acknowledgments

This work has been supported by NSERC/Discovery Grant No. 72031326. One of the authors (R.C.) is partially supported
by China Scholarship Council for his research in University of Toronto. LANL is under the auspices of US Department of En-
ergy under Contract No. DE-AC52-06NA25396.

References

[1] Ouchterlony F. On the background to the formulas and accuracy of rock fracture toughness measurements using ISRM standard core specimens. Int J
Rock Mech Min Sci Goemech Abstr 1989;26(1):13–23.
[2] Hanson JH, Ingraffea AR. Standards for fracture toughness testing of rock and manufactured ceramics: what can we learn for concrete? Cement
Concrete Aggr 1997;19(2):103–11.
[3] Fowell RJ, Xu C. The cracked chevron-notched Brazilian disc test geometrical considerations for practical rock fracture toughness measurement. Int J
Rock Mech Min Sci Goemech Abstr 1993;30(7):821–4.
[4] Chong KP, Kuruppu MD. New specimen for fracture-toughness determination for rock and other materials. Int J Fracture 1984;26(2):R59–62.
[5] Zhang ZX, Kou SQ, Yu J, Yu Y, Jiang LG, Lindqvist PA. Effects of loading rate on rock fracture. Int J Rock Mech Min 1999;36(5):597–611.
[6] Zhang ZX, Kou SQ, Jiang LG, Lindqvist PA. Effects of loading rate on rock fracture: fracture characteristics and energy partitioning. Int J Rock Mech Min
2000;37(5):745–62.
[7] Böhme W, Kalthoff JF. The behavior of notched bend specimens in impact testing. Int J Fracture 1982;20(4):R139–43.
[8] Maigre H, Rittel D. Dynamic fracture detection using the force–displacement reciprocity: application to the compact compression specimen. Int J
Fracture 1995;73(1):67–79.
[9] Owen DM, Zhuang S, Rosakis AJ, Ravichandran G. Experimental determination of dynamic crack initiation and propagation fracture toughness in thin
aluminum sheets. Int J Fracture 1998;90(1-2):153–74.
[10] Weerasooriya T, Moy P, Casem D, Cheng M, Chen W. A four-point bend technique to determine dynamic fracture toughness of ceramics. J Am Ceramic
Soc 2006;89(3):990–5.
[11] Bertram A, Kalthoff JF. Crack propagation toughness of rock for the range of low to very high crack speeds. In: Buchholz FG, Richard HA, Aliabadi MH,
editors. Advances in fracture and damage mechanics key engineering materials. Uetikon-Zurich: Trans Tech Publications; 2003. p. 423–30.
[12] Freund LB. Dynamic fracture mechanics. Cambridge: Cambridge University Press; 1990.
[13] Xia K, Chalivendra VB, Rosakis AJ. Observing ideal self-similar crack growth in experiments. Engng Fract Mech 2006;73(18):2748–55.
[14] Griffith LJ, Martin DJ. Study of dynamic behavior of a carbon–fiber composite using split Hopkinson pressure bar. J Phys D – Appl Phys
1974;7(17):2329–44.
[15] Tang CN, Xu XH. A new method for measuring dynamic fracture-toughness of rock. Engng Fract Mech 1990;35(4-5):783–9.
[16] Ramesh KT, Kelkar N. Technique for the continuous measurement of projectile velocities in plate impact experiments. Rev Sci Instrum
1995;66(4):3034–6.
[17] Ramesh KT, Narasimhan S. Finite deformations and the dynamic measurement of radial strains in compression Kolsky bar experiments. Int J Solids
Struct 1996;33(25):3723–38.
[18] Suresh S, Ewart L, Maden M, Slaughter WS, Nguyen M. Fracture-toughness measurements in ceramics – pre-cracking in cyclic compression. J Mater Sci
1987;22(4):1271–6.
[19] James MN, Human AM, Luyckx S. Fracture-toughness testing of hardmetals using compression–compression precracking. J Mater Sci
1990;25(11):4810–4.
[20] Bergmann G, Vehoff H. Precracking of nial single-crystals by compression–compression fatigue and its application to fracture-toughness testing.
Scripta Metall Mater 1994;30(8):969–74.
[21] Lim IL, Johnston IW, Choi SK, Boland JN. Fracture testing of a soft rock with semicircular specimens under 3-point bending1. Mode-I. Int J Rock Mech
Min Sci Goemech Abstr 1994;31(3):185–97.
[22] Nasseri MHB, Mohanty B. Fracture toughness anisotropy in granitic rocks. Int J Rock Mech Min 2008;45(2):167–93.
[23] Kolsky H. Stress waves in solids. Oxford: Clarendon Press; 1953.
[24] Graff KF. Wave motion in elastic solids. Oxford: Clarendon Press; 1975.
[25] Frew DJ, Forrestal MJ, Chen W. Pulse shaping techniques for testing brittle materials with a split Hopkinson pressure bar. Exp Mech
2002;42(1):93–106.
[26] Chen W, Zhang B, Forrestal MJ. A split Hopkinson bar technique for low-impedance materials. Exp Mech 1999;39(2):81–5.
[27] Frew DJ, Forrestal MJ, Chen W. Pulse shaping techniques for testing elastic–plastic materials with a split Hopkinson pressure bar. Exp Mech
2005;45(2):186–95.
[28] ASTM Standard E399-90. Standard test method for plane strain fracture toughness of metallic materials. Annual book of ASTM Standards, ASTM
International; 2002.
[29] Barsoum RS. Triangular quarter-point elements as elastic and perfectly-plastic crack tip elements. Int J Numer Meth Engng 1977;11(1):85–98.
[30] Iqbal N, Mohanty B. Experimental calibration of stress intensity factors of the ISRM suggested cracked chevron-notched Brazilian disc specimen used
for determination of mode-I fracture toughness. Int J Rock Mech Min 2006;43(8):1270–6.
[31] Song B, Chen W. Energy for specimen deformation in a split Hopkinson pressure bar experiment. Exp Mech 2006;46(3):407–10.

You might also like