You are on page 1of 13

Downloaded from http://rsta.royalsocietypublishing.

org/ on December 16, 2016

Hugoniot equation of state of


rock materials under shock
rsta.royalsocietypublishing.org
compression
Q. B. Zhang1 , C. H. Braithwaite2 and J. Zhao1
Research 1 Department of Civil Engineering, Monash University, Clayton,

Victoria 3800, Australia


Cite this article: Zhang QB, Braithwaite CH, 2 Cavendish Laboratory, JJ Thomson Avenue, Cambridge CB3 0HE, UK
Zhao J. 2017 Hugoniot equation of state of rock
materials under shock compression. Phil. QBZ, 0000-0002-1162-6557
Trans. R. Soc. A 375: 20160169.
Two sets of shock compression tests (i.e. conventional
http://dx.doi.org/10.1098/rsta.2016.0169
and reverse impact) were conducted to determine the
shock response of two rock materials using a plate
Accepted: 21 September 2016 impact facility. Embedded manganin stress gauges
were used for the measurements of longitudinal stress
and shock velocity. Photon Doppler velocimetry was
One contribution of 15 to a theme issue
used to capture the free surface velocity of the target.
Experimental testing and modelling of brittle Experimental data were obtained on a fine-grained
materials at high strain rates. marble and a coarse-grained gabbro over a shock
pressure range of approximately 1.512 GPa. Gabbro
Subject Areas: exhibited a linear Hugoniot equation of state (EOS) in
geology, civil engineering, engineering the pressureparticle velocity (Pup ) plane, while for
geology marble a nonlinear response was observed. The EOS
relations between shock velocity (US ) and particle
Keywords: velocity (up ) are linearly fitted as US = 2.62 + 3.319up
and US = 5.4 85 + 1.038up for marble and gabbro,
strain rate, shock compression, plate impact,
respectively.
equation of state, dynamic loading, rock This article is part of the themed issue Experimental
materials testing and modelling of brittle materials at high strain
rates.
Author for correspondence:
Q. B. Zhang
e-mail: qianbing.zhang@monash.edu 1. Introduction
The Hugoniot properties of rock materials have long
been a source of interest, and, traditionally, much of
the research reported in the literature is concerned
with shock properties at levels consistent with the
lower mantle [1], underground nuclear explosions [2,3],
planetary impacts [4,5] and geophysical applications [6].
For example, Ahrens [7] reviewed the equation of state
(EOS) of materials under high pressures in the range of
10400 GPa [7]. The data from the Los Alamos National
Laboratory [8,9], Trunin and co-workers [1012] and
Stffler [13,14] are again generally at higher pressures.

2016 The Author(s) Published by the Royal Society. All rights reserved.
Downloaded from http://rsta.royalsocietypublishing.org/ on December 16, 2016

Table 1. Summary of selected shock compression data from plate impact experiments in the literature.
2
elastic

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 375: 20160169


.........................................................
modulus E
rock type (g cm3 ) C L (km s1 ) (GPa) V p (m s1 ) US (m s1 ) x (GPa) references
Bukit Timah granite 2.67 5.82 75.2 65.1801.3 30.9369.3 0.4966.99 [30]
..........................................................................................................................................................................................................

Swedish gabbro 2.88 6.21 n.a. 521862 355674 6.111.9 [31]


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

Loughborough granite 2.65 5.13 n.a. 368905 200768 2.711.1 [31]


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

dolerite 2.89 5.89 n.a. 451833 310670 5.1611.34 [33]


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

kimberlite breccia 2.49 3.56 31.9 0.8 175900 79680 0.348.4 [19]
..........................................................................................................................................................................................................

Recently, there has been a growing interest in the shock properties of rock materials subjected
to mining explosive and blasting loadings [1522]. There are important differences between the
high-pressure studies and shock Hugoniot at the lower pressures that are relevant to mining
situations. Above 20 GPa, shock waves in rock materials are largely hydrodynamic and the effects
of material strength and structure can generally be neglected [19]. However, elastic and elastic
plastic behaviours and the effects of material structure should be considered at low pressures
(below 20 GPa) [23]. Experimental data are not comprehensive in the pressure range lower
than 20 GPa, although some examples are quartz, calcite and plagioclase rocks [24], Westerly
granite and Nugget sandstone [25], anorthosite and gabbro [26], Hunters Trophy tuff, UTTR
limestone, Pennsylvania slate and permafrost phyllite [27], Kinosaki basalt [28,29], Bukit Timah
granite [30], Swedish gabbro and Loughborough granite [31], Berea sandstone [32], dolerite [33],
limestone [34], kimberlite breccia [19] and Westerly granite [35].
The methods and devices most widely used to apply shock loading to specimens are in-contact
explosives, explosively driven flyer plates, pulsed-radiation devices and gun-type launchers [36
39]. It is important to recognize that there is often significant scatter on the measurements due
to the lack of standard experimental techniques and procedures. Gas guns in a plate impact
configuration have the advantage of being able to be in non-explosive controlled areas and have
good control over projectile velocity and therefore pressure induced in the samples. Table 1
summarizes selected shock compression data of rock materials at pressure levels up to 20 GPa
from plate impact experiments in the literature. The motivation of this work is to perform
standard experimental procedures for the determination of dynamic shock responses of rock
materials using a plate impact facility.

2. Equation of state: Hugoniot properties


A shock wave can be defined as a traveling wave front across which a discontinuous adiabatic
jump in state variables occurs [40, p. 946]. There are several excellent reviews and books on shock
compression behaviour of materials [7,10,37,4146]. From the point of view of a surfer travelling
with the boundary, the mass moving towards the front and the mass moving away from the
front per unit area must equal [37,42]. For planar impact between two materials, the simplified
RankineHugoniot jump conditions, which describe the conservation of mass, momentum and
energy across a steady shock wave, are expressed as follows:

0 US = 1 (US up ), (2.1)
P1 P0 = 0 (US u0 )(up u0 ) (2.2)

and e1 e0 = 12 (P1 + P0 )(V1 V0 ), (2.3)

where the subscripts 0 and 1 refer to the states of unshocked and shocked material, respectively.
P, e, V and describe the pressure, energy, volume and density, V = 1/ is the specific volume,
US is the shock velocity and up is the particle velocity.
Downloaded from http://rsta.royalsocietypublishing.org/ on December 16, 2016

Equations (2.1)(2.3) are derived assuming no heat conduction and are commonly referred to
3
as the RankineHugoniot relations. In equation (2.2), when P0 , usually equal to 1 atmosphere, is
considered negligibly small compared with P1 , if u0 = 0,

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 375: 20160169


.........................................................
P1 = 0 US up , (2.4)

where 0 US is often called the shock impedance.


In the above conservation equations, there are five parameters (P, V, e, Us and up ). Thus, an
additional equation is required if one needs to determine all parameters as a function of one
of them. One way of determining this experimentally is to measure the shock Hugoniot of the
material, giving the shock equation of state. This equation can be frequently expressed as the
relationship between shock velocity and particle velocity. For most of materials, the Hugoniot
EOS is represented by a linear relation [42],

US = C0 + Sup . (2.5)

The constants C0 and S are generally determined by experiments, and the first is related to the
bulk wave speed.

3. Experimental procedures
(a) Material characteristics
Two types of well-investigated rock material were chosen for the experiments, namely one
metamorphic rock (Carrara marble) and one igneous rock (South Africa gabbro) [21], which are
also frequent in the Earths crust. Figure 1 shows the prepared square plates and thin sections
under cross-polarized light. Carrara marble has a nearly pure calcite composition (98% calcite,
and traces of dolomite, mica and quartz), and the average grain size is approximately 0.15 mm.
The gabbro is composed of 55% plagioclase, 35% orthopyroxene, 5% quartz, 3% olivine and
2% biotite, and the average grain size is around 1.50 mm. These two rocks do not have shape-
preferred or crystallographic-preferred orientations. The longitudinal and shear wave speeds (CL
and CS ) were obtained by a time of flight method using a 5 MHz ultrasonic transducer. During
shock wave experiments, various standard materials were used as flyer plates, target plates and
windows for photon Doppler velocimetry (PDV) measurements. In this study, the materials are
polymethyl methacrylate (PMMA), aluminium alloy HE30 (which is essentially equivalent to
Al6082-T6), and copper C101. Extensive acoustic sound velocities and shock wave experiments
have been conducted to characterize these materials by Chapman [47]. The physical properties
and Hugoniot data of each material are summarized in table 2.

(b) The plate impact facility


Shock compression experiments were performed using the plate impact facility at the Cavendish
Laboratory, Cambridge, UK [48], as schematically depicted in figure 2. The facility consists of a
single-stage light gas gun with a 5 cm bore and a 5 m barrel length capable of launching projectiles
at up to 1100 m s1 . In the target chamber, the specimen mount can be adjusted by means of three
screws (generally achieving a tilt of approx. 1 mrad or better), and a variety of diagnostic and
soft-recovery techniques can be applied.
Deriving data from plate impact experiments relies on measuring, either directly or indirectly,
two of the five parameters (P, V, e, US and up ) necessary for a full description of the material
response to shock loading. There are several variables of interest in the plate impact experiments,
and excellent summaries of measurement techniques are given by Bourne [37] and Field et al. [44].
The research presented here involves the measurements of the impact velocity of the projectile
(V P ), the longitudinal stresses ( x ), the shock velocity (US ) and the particle velocity (up ) as a
function of time.
Downloaded from http://rsta.royalsocietypublishing.org/ on December 16, 2016

(a) (b) 4
(i) (i)

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 375: 20160169


.........................................................
10 mm 10 mm

(ii) (ii)

2 mm 2 mm

Figure 1. Photographs of the prepared specimens ((i) with a 10 mm scale) and the thin sections under cross-polarized light
((ii) with a 2 mm scale) of: (a) Carrara marble and (b) South Africa gabbro. (Online version in colour.)

Table 2. Physical properties and the Hugoniot data of rock and standard materials.

Poissons
material type (g cm3 ) C L (km s1 ) C S (km s1 ) C 0 (km s1 ) S E (GPa) ratio reference
marble 2.680 5.340 2.800 4.250 40 0.20
..........................................................................................................................................................................................................

gabbro 2.900 6.500 3.020 5.485 61 0.13


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

PMMA 1.187 2.748 1.392 2.598 1.516 [47]


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

Al6082-T6 2.703 6.418 3.122 5.380 1.337


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

copper C101 8.930 4.781 2.335 3.940 1.489


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

A series of shorting pins were applied to determine the impact velocity of the projectile.
The embedded manganin stress gauges (LM-SS-125CH-048, approx. 50 m thick; Vishay Micro-
Measurements) give a direct output of the longitudinal stress in the material of interest. Shock
velocity is the easiest parameter to measure, from the time of arrival between two longitudinal
gauges at a known distance apart. The voltage output from the gauges is recorded on an
oscilloscope (Tektronix TDS 7054) at a sample rate of up to 5 GS s1 . PDV has become one of the
Downloaded from http://rsta.royalsocietypublishing.org/ on December 16, 2016

vacuum chamber
5
oscilloscopes
flyer plate target assembly
breech , ufs
50 mm single-stage gas gun

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 375: 20160169


.........................................................
high-pressure
gas tank
polycarbonate VP PDV probe
projectile
stress gauge trigger signal

velocity target holder


vacuum pump pins oscilloscope
VP

Figure 2. Schematic of the plate impact facility at the Cavendish Laboratory with measurement systems. (Online version in
colour.)

moving target
v(t) f0, incident light
fd
f0, reference light
cleaved fibre
f0 fd, signal light

Miteq Tektronix
Thorlabs
2 DR-125G-A TPO7254
SFL1550S f0
1 3
laser detector oscilloscope

40 mW, 1550 nm circulator 13 GHz 2.5 GHz, 40 GS s1


Thorlabs 6015-3-APC

Figure 3. Schematic view of the PDV. (Online version in colour.)

most commonly used interferometer systems for measuring the free surface velocity of targets
and works by measuring a beat frequency between a reference source and light Doppler shifted
from the target [49],

I(t) = I0 + Id + I0 Id sin[fb (t) + ], (3.1)
fb (t) = |fd (t) f0 | (3.2)

and v(t) = fb (t), (3.3)
2

where I0 is the non-Doppler-shifted intensity from the laser, Id is the Doppler-shifted intensity
from the moving surface, f b is the beat frequency, is the relative phase between the Doppler-
shifted and non-Doppler-shifted light, f d (t) is the Doppler-shifted frequency, f 0 is the reference
frequency of the laser, and is the wavelength of the continuous wave (CW) laser. The beat
frequency f b (t) is related to the velocity v(t).
As schematically shown in figure 3, the PDV system at the Cavendish Laboratory [50] has
a high-power 1550 nm CW distributed feedback laser with a polarization-maintaining fibre
(Thorlabs SFL1550S). The laser is operated at a maximum power of 40 mW with a linewidth
of 50 kHz. A photodiode detector (Miteq DR-125G-A) with bandwidth 12.5 GHz is employed.
Downloaded from http://rsta.royalsocietypublishing.org/ on December 16, 2016

(a) (b)
6
flyer Mylar sheet Mylar sheet
P target

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 375: 20160169


.........................................................
P1 1 rear surface plate

P2 2
reflected target

rock plate rock plate


0 0 5 mm PMMA
SG1 10 mm SG2
up1 Vp up 12 mm

(c) (d)

Figure 4. Schematic of the conventional compression test (a) showing data derivation with a rear surface gauge (b) (exploded)
of a target enclosed within two longitudinal stress gauge (SG) packages. (c) The sabot and the copper flyer and (d) the fixed
Mylar sheet and stress gauges. (Online version in colour.)

The heterodyne signals are recorded with an oscilloscope (Tektronix DPO7254) operating with a
bandwidth of 2.5 GHz and a sampling rate of up to 40 GS s1 .

(c) Testing methods


The specific flyer and specimen geometry, and the locations of the sensors within the specimen,
depend on both the investigated materials and the properties being measured. The diameter of
the target plate is typically chosen to be greater than that of the flyer disc. All of the specimens
were ground flat and parallel to a tolerance of approximately 50 m to ensure planar shock wave
conditions. Target specimens were typically constructed from a number of sample plates. As the
flyer and target plates are of finite dimension, radial release waves will eventually converge if
the axis is 45 [51], and longitudinal releases will also eventually overtake the shock, relieving
the pressure. Thus to measure the steady shock state, the flyer thickness and the sensor locations
should be chosen such that the release waves do not interfere with the shock during the period of
interest [47].
The conventional impact configuration, as shown in figure 4, allows for the measurement of
longitudinal stress and shock velocity using longitudinal gauges. To insulate and protect the
gauges, a 25 m thick Mylar sheet was fixed to the specimen plates using a slow cure epoxy
(Loctite 0151 Hysol Epoxy Adhesive). The plate and Mylar sheet were clamped using a
specially machined clamp apparatus for a minimum of 24 h until the slow cure epoxy set. The
gauge was glued on one specimen plate with a drop of instant adhesive (Loctite 406 Prism),
then the other plate was fixed on the surface of the plate with the gauge using a small amount of
epoxy. The rear surface gauge employed a 12 mm thick PMMA backing plate to provide support
for the gauge. All components, glued together using the epoxy, were again clamped until the
epoxy set. The gauge package, comprising the gauge, Mylar sheet and the epoxy, is typically about
150 50 m thick. The impedance difference between the gauge package and the target causes
the stress in the gauge to rise to the stress within the target through repeated boundary reflections
in about 200300 ns. In the conventional configuration, figure 4a gives graphical representations
Downloaded from http://rsta.royalsocietypublishing.org/ on December 16, 2016

(a) t (b) P rock Hugoniot 7


rock copper 1
rock release

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 375: 20160169


.........................................................
copper loading
copper unloading
8
7
3
VP 6
5 PDV
5
4
3
7
2
1 2 4 6 8
X up

(c)

Al ring

marble

Figure 5. (a) Schematic of the reverse impact experiment in an Xt diagram showing shocks and releases in the copper plate.
(b) The stressparticle velocity space showing the determination of a release path. (c) The flyer (a specially machined rock disc
and an aluminium ring) and a well-prepared sabot. (Online version in colour.)

of the associated shock states in the stressparticle velocity space. The initial impact puts the
rock between the two gauges in a state represented by the crossing point of the flyer and
target Hugoniot. As the shock interacts with the rear surface plate (PMMA), a release (since the
impedance of PMMA is lower than that of rock) propagates back into the rock and the rear surface
material is shocked up to the state marked as 2. However, it should be noted that this assumes that
the release in the rock is the reverse of the Hugoniot. Because of the similar impedance between
the gauge package and PMMA, the measured stress in the PMMA ( PMMA ) can be converted to
stress in the rock target ( T ) through a well-known relation [31]

ZT + ZPMMA
T = PMMA , (3.4)
2ZPMMA
where ZT is the shock impedance (ZT = 0 US ) of the target and ZPMMA is the elastic impedance
of PMMA.
The reverse configuration of shock compression is shown in figure 5. A 48 mm diameter rock
flyer with a thickness of 10 mm was impacted onto a copper target. For non-conducting rock
materials, an aluminium ring is added on the front, as shown in figure 5c, so that the projectile
velocity measurement system works correctly. The PDV system was used to measure the surface
particle velocity and allow for the release properties of the flyer to be investigated. Figure 5a
shows the Lagrangian distancetime (Xt) diagram of the experiment. The frame of reference is
the interface between the flyer and the target. Initially, a shock wave travels back into the flyer
and forward into the target plate. These waves are both reflected as releases from the free surfaces.
The release from the rear of the flyer, due to the relative thickness of the flyer and the target plate,
is irrelevant to the remainder of the experiment. The wave in the copper plate then reflects back
and forth within the plate as a series of shocks and releases. As there is continuity of pressure and
Downloaded from http://rsta.royalsocietypublishing.org/ on December 16, 2016

particle velocity at the rockcopper interface, the reloading of the copper is characteristic of the
8
release state in the rock, as illustrated in figure 5b.

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 375: 20160169


.........................................................
4. Results and discussion
The Hugoniot elastic limit (HEL) is the threshold on the shock Hugoniot at which a material
transitions from an elastic state to an elasticplastic state. HEL values of polycrystalline ceramics
were given by Kanel et al. [16], but the HEL of rock materials is seldom reported or is detected only
in a few shots in the literature [52]. One common way to determine an HEL value is to observe the
amplitudes of the elastic precursor waves (i.e. two-wave structures) in diagnostic records [24,53].
Figure 6a shows a typical voltagetime trace of gabbro measured by two embedded longitudinal
gauges at the impact velocity V P of 826 m s1 , which is the highest in this study. The traces rise
from the initial ambient level to a peak value, but only one wave structure is observed, as also
reported by many researchers even at the impact velocity V P up to 3490 m s1 [28], perhaps
due to the similar values of shock impedance and elastic impedance [31,33,54]. The longitudinal
stress x is calculated from the voltagetime trace [55], and the shock velocity US is determined
from the time of arrival of shock waves at two gauge locations with the known distance. The
particle velocity up can be calculated from equation (2.4) ( = P1 = 0 US up ). Figure 4a illustrates
the determination of the Hugoniot parameters using the slope of the target Rayleigh line 0 US .
In addition, the Hugoniot parameters can also be calibrated by x , V P and the known EOS of the
standard flyer material.
In the reverse impact tests, the beat waveform in figure 6b shows amplitude fluctuations of the
moving surface, and the expanded time base in figure 6c reveals the individual beat cycles. The
surface of the target moves through a distance equal to one-half of the laser wavelength (1550 nm),
and a measured period of 5.1 ns corresponds to a free surface velocity ufs of 152 m s1 . Figure 6d
shows the free surface velocity trace from the spectrogram calculated by Fourier transform
and the numbers (2, 4, 6 and 8) denote the shock states for the determination of the shock
release behaviour. The Hugoniot state can be determined from the slope of the target Rayleigh
line 0 (C0 + Sup ), and the particle velocity up (approx. up = 0.5ufs , as described by Duvall &
Fowles [56]) and impact velocity V P , as schematically shown in figure 5b.
While each of the methods should lead to the same Hugoniot data, it should be mentioned
that the method has advantages and disadvantages depending on precisely the measurement
and the material being investigated. The PDV system has excellent time resolution, but monitors
only the material surface. Manganin gauges provide in-material results, and simultaneously with
the use of more than one gauge to determine the opportunity to measure shock velocity and/or
shock wave attenuation. In addition, the cost of the gauges is significantly less than that of the
PDV system. However, it has also been demonstrated that for certain rock-like materials the noise
associated with gauge traces is sufficient to warrant using a reverse impact configuration and
the PDV system to obtain more reliable results [57]. Shock behaviour plotted in Pup space is
preferred in this study because either P or up is directly measured in experiments. The other
parameters in US up and PV graphs can be computed by classical shock equations.
Figure 7a presents the measured Hugoniot states for these rock materials in shock velocity
particle velocity (US up ) space. The EOS relations between US and up are linearly fitted
US = 5.485 + 1.038up and US = 2.62 + 3.319up for gabbro and marble, respectively. Although US
up results of gabbro [26,31] and marble [58] were also presented in the literature, the values of
C0 and S and the bulk wave speed of rocks were not given. For gabbro, the fitted values of C0
and S are 6.409 (0.286 < up < 0.60) [26], 5.965 (0.286 < up < 0.60) [31], and 0.215 [26], 0.411 [31],
respectively. For marble, the fitted values of C0 and S are 2.57 and 2.50 (0.115 < up < 0.958) [58],
respectively. It was found that, for gabbro, the constant C0 that is the same as the bulk wave
speed of 5.485 km s1 and S = 1.038 are similar to the values for ceramics [45]. By contrast,
marble appears to have the value of C0 (2.62 km s1 ), significantly below the calculated bulk
wave speed (4.25 km s1 ) in table 2. It perhaps indicates a significant loss of strength or a phase
transition of calcite (Carrara marble has 98% calcite composition) at a pressure of approximately
Downloaded from http://rsta.royalsocietypublishing.org/ on December 16, 2016

(a) (b)
1.6 0.6 9
gabbro, VP = 826 m s1 0.10 400
marble, VP = 350 m s1
1.4 front SG
0.5 0.08 measurement 350
back SG

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 375: 20160169


.........................................................
free surface velocity (m s1)
1.2 ufs
electrical potential (V)

electrical potential (V)

electrical potential (V)


0.06 300
0.4
1.0
0.04 250
0.8 0.3
0.02 200
0.6
0.2 0 150
0.4
0.1 0.02 100
0.2
0.04 50
0 0
0.06 0
51 52 53 54 55 56 54 55 56 57 58 59 60
time (s) time (s)

(c) (d)
350 8
0.04
6
v(t) = 775/5.1 = 152 m s1 5.1 ns
300
4
0.02 250
electrical potential (V)

ufs (m s1)
200
2
0
150

100
0.02

50

0.04
55.56 55.57 55.58 55.59 55.60 55.61 55 56 57 58 59 60 61 62 63
time (s) t (ms)

Figure 6. (a) Stress gauge trace from gabbro at a V P of 826 m s1 . (b) The beat waveform showing amplitude fluctuations of
the moving surface. (c) The expanded time base revealing the individual beat cycles. (d) The velocity trace from the spectrogram
calculated by Fourier transform. (Online version in colour.)

1.2 GPa [27,45]. Moreover, marble indicates a much higher value of S, which probably reveals the
subsequent yield or transition attested by the third wave during shock compression [45].
Above the HEL, the material loses much of its shear strength, though generally this is without
the two-wave behaviours generally seen in shocked metals, for example. Attempts have been
made to measure the shear strength and to then determine the deviation from elastic behaviour
and infer an HEL [33]. In addition, the following relationship between the HEL and spall strength
( spall ) with the Griffith criterion is given by [59]
1
HEL = 8spall . (4.1)
(1 2)2
Spall strength and Poissons ratio v of most rock materials are 0.010.13 GPa [60] and 0.10
0.35 [61], respectively, suggesting that the HEL values are in the range of 0.117.51 GPa.
The Hugoniot data derived from the stress gauge and PDV records are shown in figure 7b. The
Hugoniots lie slightly below the theoretical elastic lines (P = 0 CL up ). The Hugoniot of gabbro in
the pressureparticle velocity (Pup ) plane is well fitted by a linear function P = 17.30up . The data
for marble fit better with a second-order polynomial relation P = 7.511up + 7.92u2p , which might
be induced by the phase transition of calcite [27,45].
The specific volumes of rocks are derived using the Hugoniot equations (equation (2.2),
conservation of energy), and the Hugoniot curves in pressure (P) and relative specific volume
(V 1 /V 0 ) are shown in figure 7c. Unlike the gabbro that is the linear relation, P = 96.2(1 V1 /V0 ),
the marble has a nonlinear PV curve, P = 363.5(V1 /V0 ) 704.4(V1 /V0 )2 + 340.9(V1 /V0 )3 , and
does not converge towards any value in this stress range. In addition, at a pressure close to 10 GPa,
the gabbro is compressed to approximately 90% of its initial volume, which is less than 84% of
that observed in the marble. As proposed by Eakins & Thadhani [41] the complete collapse of
Downloaded from http://rsta.royalsocietypublishing.org/ on December 16, 2016

(a) 7 (b) 12
Us = 5.485 + 1.038up (R2 = 0.87) gabbro 10
gabbro P = 17.30up
10 elastic P = 18.85up

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 375: 20160169


6

.........................................................
C0 = 5.485 8
Us (km s1)

P (GPa)
5
Us = 2.62 + 3.319up (R2 = 0.97) 6
C0 = 4.250
4 4
marble
marble
gabbro 2 marble P = 7.51up + 7.92u2p
3 marble EOS
gabbro EOS elastic P = 14.31up
0
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8
up (km s1) up (km s1)

(c) 12
marble
gabbro
10 marble P/(V1/V0)
gabbro P/(V1/V0)
8
P (GPa)

P = 96.2 (1 V1/V0)
6

2
P = 363.5(V1/V0) 704.4(V1/V0)2 + 340.9(V1/V0)3
0
0.84 0.88 0.92 0.96 1.00
V1/V0

Figure 7. Measured Hugoniot states in: (a) the shock velocityparticle velocity (US up ) space, (b) the stressparticle velocity
(Pup ) space (the elastic response is calculated by P = 0 C L up ) and (c) the pressure-specific volume (PV) space. (Online
version in colour.)

porosities occurs at high pressure. The phase transition of calcite is at a pressure of 1.2 GPa [27,45],
which is lower than the shock pressure in this study. Thus, the primary reasons are that the
gabbro with large well-formed interlocking crystals has less porosity (here, the porosity is defined
by microcracks along crystals and scattered intergranular pores) than the marble and the phase
transition of calcite in marble.

5. Conclusion
Plate impact experiments were conducted to determine the Hugoniot EOS of fine-grained marble
and coarse-grained gabbro at high pressures up to 12 GPa. Manganin stress gauges and a PDV
system were used for the measurement of stress, shock particle velocity and free surface velocity.
Two compression testing methods conducted to obtain Hugoniot curves provided satisfactory
results. The shock velocityparticle velocity (US up ) data of two rocks were fitted by straight lines
in this pressure range. The Hugoniot curves in the plane of both the pressureparticle velocity
(Pup ) and pressurerelative specific volume (P V 1 /V 0 ) are well fitted by a linear function, and
a polynomial function for gabbro and marble, respectively.
Authors contributions. Q.B.Z. made substantial contributions to the conception, acquisition of data, analysis and
interpretation of data; and drafted the article for important intellectual content. C.H.B. made substantial
contributions to the conception and design. J.Z. gave final approval of the version to be published.
Competing interests. We have no competing interests.
Downloaded from http://rsta.royalsocietypublishing.org/ on December 16, 2016

Funding. This work is supported by the Swiss National Science Foundation (grant no. 200020_129757).
11
Acknowledgements. Q.B.Z. would like to thank Prof. John Field and Dr Andrew Jardine for providing the
opportunity to study at the SMF group of the Cavendish Laboratory.

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 375: 20160169


.........................................................
References
1. Boehler R. 2000 High-pressure experiments and the phase diagram of lower mantle and core
materials. Rev. Geophys. 38, 221245. (doi:10.1029/1998RG000053)
2. Short NM. 1966 Effects of shock pressures from a nuclear explosion on mechanical and optical
properties of granodiorite. J. Geophys. Res. 71, 11951215. (doi:10.1029/JZ071i004p01195)
3. Trunin RF. 1994 Shock compressibility of condensed materials in strong shock waves
generated by underground nuclear explosions. Phys. Usp. 37, 11231145. (doi:10.1070/PU1994
v037n11ABEH000055)
4. Ross M, Graboske HC, Nellis WJ. 1981 Equation of state experiments and theory relevant to
planetary modelling. Phil. Trans. R. Soc. Lond. A 303, 303313. (doi:10.1098/rsta.1981.0204)
5. Kenkmann T, Poelchau MH, Wulf G. 2014 Structural geology of impact craters. J. Struct. Geol.
62, 156182. (doi:10.1016/j.jsg.2014.01.015)
6. Anderson OL. 1995 Equations of state of solids for geophysics and ceramic science. Oxford, UK:
Oxford University Press.
7. Ahrens TJ. 1993 Equation of state. In High-pressure shock compression of solids (eds JR Asay, M
Shahinpoor), pp. 75113. New York, NY: Springer.
8. Marsh SP. 1980 LASL shock Hungoniot data. Berkeley, CA: University of California Press.
9. McQueen RG, Marsh SP, Fritz JN. 1967 Hugoniot equation of state of twelve rocks. J. Geophys.
Res. 72, 49995036. (doi:10.1029/JZ072i020p04999)
10. Trunin R. 1998 Shock compression of condensed materials. Cambridge, UK: Cambridge University
Press.
11. Trunin R, Simakov G, Podurets M, Moiseyev B, Popov L. 1971 Dynamic compressibility of
quartz and quartzite at high pressure. Izv. Acad. Sci. USSR Phys. Solid Earth 1, 1320.
12. Trunin RF. 2001 Shock compression of condensed materials (laboratory studies). Phys. Usp.
44, 371396. (doi:10.1070/PU2001v044n04ABEH000919)
13. Stffler D. 1982 Terrestrial impact breccias. In Lunar breccias and soils and their meteoritic
analogs (eds GJ Taylor, LL Wilkening), pp. 139146. LPI Technical Report 82-02.
Houston, TX: Lunar and Planetary Institute. See http://articles.adsabs.harvard.edu//
full/1982lbsm.work..139S/0000146.000.html.
14. Stffler D. 1982 1.3 Density of minerals and rocks under shock compression. In Landolt-
Brnsteingroup V geophysics 1A (subvolume A) (eds G Angenheister), pp. 120126. Berlin,
Germany: Springer.
15. Grady DE, Kipp ME. 1979 The micromechanics of impact fracture of rock. Int. J. Rock Mech.
Min. Sci. Geomech. Abstr. 16, 293302. (doi:10.1016/0148-9062(79)90240-7)
16. Kanel GI, Bless SJ, Rajendran AM. 2000 Behavior of brittle materials under dynamic loading.
Technical report no. ADA386439. Institute for Advanced Technology, The University of Texas
at Austin, Austin, TX, USA. See http://www.dtic.mil/cgi-bin/GetTRDoc?AD=ADA386439.
17. Rossmanith HP, Daehnke A, Nasmillner REK, Kouzniak N, Ohtsu M, Uenishi K. 1997 Fracture
mechanics applications to drilling and blasting. Fatig. Fract. Eng. Mater. Struct. 20, 16171636.
(doi:10.1111/j.1460-2695.1997.tb01515.x)
18. Zhao J et al. 1999 Rock dynamics research related to cavern development for ammunition
storage. Tunn. Undergr. Sp. Tech. 14, 513526. (doi:10.1016/S0886-7798(00)00013-4)
19. Willmott GR, Proud WG. 2007 The shock Hugoniot of Tuffisitic Kimberlite Breccia. Int. J. Rock
Mech. Min. 44, 228237. (doi:10.1016/j.ijrmms.2006.07.006)
20. Braithwaite C. 2009 High strain rate properties of geological materials. Cambridge, UK: University
of Cambridge.
21. Zhang QB. 2014 Mechanical behaviour of rock materials under dynamic loading. Lausanne,
Switzerland: cole Polytechnique Fdrale de Lausanne.
22. Kirk S. 2014 Shock compression and dynamic fragmentation of geological materials. Cambridge, UK:
University of Cambridge.
Downloaded from http://rsta.royalsocietypublishing.org/ on December 16, 2016

23. Lysne PC. 1970 A comparison of calculated and measured low-stress Hugoniots
12
and release adiabats of dry and water-saturated tuff. J. Geophys. Res. 75, 43754386.
(doi:10.1029/JB075i023p04375)

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 375: 20160169


.........................................................
24. Ahrens TJ, Gregson Jr VG. 1964 Shock compression of crustal rocks: data for quartz, calcite,
and plagioclase rocks. J. Geophys. Res. 69, 48394874. (doi:10.1029/JZ069i022p04839)
25. Larson DB, Anderson GD. 1980 Plane shock wave studies of Westerly granite and
Nugget sandstone. Int. J. Rock Mech. Min. Sci. Geomech. Abstr. 17, 357363. (doi:10.1016/
0148-9062(80)90519-7)
26. Boslough MB, Ahrens TJ. 1985 Shock wave properties of anorthosite and gabbro. J. Geophys.
Res. 90, 78147820. (doi:10.1029/JB090iB09p07814)
27. Davies FW, Smith EA. 1994 High pressure equation of state investigation of rocks. Contract no.
ADA284761. Albuquerque, NM: Ktech Corp.
28. Nakazawa S, Watanabe S, Kato M, Iijima Y, Kobayashi T, Sekine T. 1997 Hugoniot equation
of state of basalt. Planet. Space Sci. 45, 14891492. (doi:10.1016/S0032-0633(97)00070-6)
29. Sekine T, Kobayashi T, Nishio M, Takahashi E. 2008 Shock equation of state of basalt. Earth
Planets Space 60, 9991003. (doi:10.1186/BF03352857)
30. Shang JL, Shen LT, Zhao J. 2003 Hugoniot equation of state of the Bukit Timah granite. Int. J.
Rock Mech. Min. 37, 705713. (doi:10.1016/S1365-1609(00)00002-2)
31. Millett JCF, Tsembelis K, Bourne NK. 2000 Longitudinal and lateral stress measurements in
shock-loaded gabbro and granite. J. Appl. Phys. 87, 36783682. (doi:10.1063/1.372399)
32. Lomov IN, Hiltl M, Vorobiev OY, Glenn LA. 2001 Dynamic behavior of berea sandstone
for dry and water-saturated conditions. Int. J. Impact Eng. 26, 465474. (doi:10.1016/
S0734-743X(01)00097-5)
33. Tsembelis K, Proud WG, Field JE. 2002 The principal Hugoniot and dynamic strength of
Dolerite under shock compression. AIP Conf. Proc. 620, 13851388. (doi:10.1063/1.1483797)
34. Johnson D, Chapman DJ, Tsembelis K, Proud WG. 2007 The response of dry limestone to
shock-loading. AIP Conf. Proc. 995, 13871390. (doi:10.1063/1.2832984)
35. Yuan F, Prakash V. 2013 Plate impact experiments to investigate shock-induced inelasticity in
Westerly granite. Int. J. Rock Mech. Min. 60, 277287. (doi:10.1016/j.ijrmms.2012.12.024)
36. Zhernokletov MV. 2006 Methods and devices for producing intense shock loads. In Material
properties under intensive dynamic loading. Shock wave and high pressure phenomena (eds MV
Zhernokletov, BL Glushak), pp. 3371. Berlin, Germany: Springer.
37. Bourne N. 2013 Materials in mechanical extremes: fundamentals and applications. Cambridge, UK:
Cambridge University Press.
38. Ahrens TJ. 1987 6. Shock wave techniques for geophysics and planetary physics. In Methods
in experimental physics, vol. 24, part A (eds GS Charles, LH Thomas), pp. 185235. New York,
NY: Academic Press.
39. Lexow B, Wickert M, Thoma K, Schfer F, Poelchau M, Kenkmann T. 2013 The extra-large
light-gas gun of the Fraunhofer EMI: applications for impact cratering research. Meteorit.
Planet. Sci. 48, 37. (doi:10.1111/j.1945-5100.2012.01427.x)
40. Ramesh KT. 2008 High rates and impact experiments. In Springer handbook of experimental solid
mechanics (ed. WN Sharpe), pp. 929960. New York, NY: Springer.
41. Eakins DE, Thadhani NN. 2009 Shock compression of reactive powder mixtures. Int. Mater.
Rev. 54, 181213. (doi:10.1179/174328009X461050)
42. Meyers MA. 1994 Dynamic behavior of materials. New York, NY: John Wiley & Sons, Inc.
43. Bourne NK, Millett JCF, Gray III GT. 2009 On the shock compression of polycrystalline metals.
J. Mater. Sci. 44, 33193343. (doi:10.1007/s10853-009-3394-y)
44. Field JE, Walley SM, Proud WG, Goldrein HT, Siviour CR. 2004 Review of experimental
techniques for high rate deformation and shock studies. Int. J. Impact Eng. 30, 725775.
(doi:10.1016/j.ijimpeng.2004.03.005)
45. Grady DE. 1998 Shock-wave compression of brittle solids. Mech. Mater. 29, 181203.
(doi:10.1016/S0167-6636(98)00015-5)
46. Hiermaier S. 2008 Structures under crash and impact: continuum mechanics, discretization and
experimental characterization. Berlin, Germany: Springer Science & Business Media.
47. Chapman DJ. 2009 Shock compression of porous materials and diagnostic development. Cambridge,
UK: University of Cambridge.
Downloaded from http://rsta.royalsocietypublishing.org/ on December 16, 2016

48. Bourne NK, Rosenberg Z, Johnson DJ, Field JE, Timbs AE, Flaxman RP. 1995 Design
13
and construction of the UK plate impact facility. Meas. Sci. Technol. 6, 1462. (doi:10.1088/
0957-0233/6/10/005)

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 375: 20160169


.........................................................
49. Strand OT, Goosman DR, Martinez C, Whitworth TL, Kuhlow WW. 2006 Compact system
for high-speed velocimetry using heterodyne techniques. Rev. Sci. Instrum. 77, 083108.
(doi:10.1063/1.2336749)
50. Lea LJ, Jardine AP. 2016 Application of photon Doppler velocimetry to direct impact
Hopkinson pressure bars. Rev. Sci. Instrum. 87, 023101. (doi:10.1063/1.4940935)
51. Altshuler LV, Kormer SB, Brazhnik MI, Vladimirov LA, Speranskaya MP, Funtikov AI. 1960
The isentropic compressibility of aluminium, copper, lead, and iron at high pressures. [English
translation.] Sov. Phys. JETP. 11, 761775. See http://www.jetp.ac.ru/cgi-bin/dn/e_011_04_
0766.pdf.
52. Zhang QB. 2016 Shear strength and Hugoniot elastic limit (HEL) of rock materials under shock
loading. In Rock dynamics: from research to engineering (eds HB Li, JC Li, QB Zhang, J Zhao),
pp. 227232. Boca Raton, FL: CRC Press .
53. Murri WJ, Grady DE, Mahrer KD. 1975 Equation of state of rocks. SRI Project PYU-1883.
Stanford Research Institute, Menlo Park, CA, USA.
54. Millett JCF, Tsembelis K, Bourne NK, Field JE. 2000 The shock Hugoniot of two igneous rocks.
AIP Conf. Proc. 505, 12471250. (doi:10.1063/1.1303687)
55. Rosenberg Z, Yaziv D, Partom Y. 1980 Calibration of foil-like manganin gauges in planar
shock wave experiments. J. Appl. Phys. 51, 37023705. (doi:10.1063/1.328155)
56. Duvall G, Fowles G. 1963 Shock waves. In High pressure physics and chemistry, vol. 2 (ed. RS
Bradley), pp. 209291. New York, NY: Academic Press.
57. Guest AR, Braithwaite CH, Proud WG, Field JE. 2007 The shock Hugoniot properties of
geological materials and relationship to static properties. AIP Conf. Proc. 995, 13791382.
(doi:10.1063/1.2832981)
58. Furnish M. 1994 Dynamic properties of Indiana, Fort Knox and Utah test range limestones and Danby
marble over the stress range 1 to 20 GPa. Albuquerque, NM: Sandia National Labs.
59. Rosenberg Z. 1993 On the relation between the Hugoniot elastic limit and the yield strength
of brittle materials. J. Appl. Phys. 74, 752753. (doi:10.1063/1.355247)
60. Grady DE, Hollenbach RE. 1979 Dynamic fracture strength of rock. Geophys. Res. Lett. 6, 5.
(doi:10.1029/GL006i002p00073)
61. Gercek H. 2007 Poissons ratio values for rocks. Int. J. Rock Mech. Min. Sci. 44, 113.
(doi:10.1016/j.ijrmms.2006.04.011)

You might also like