Fra FTG

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Submitted to Fatigue & Fracture of Engineering Materials, June 1999

DYNAMIC STRENGTH OF MATERIALS

G.I. Kanel
High Energy Density Research Centre of Russian Academy of Sciences,
IVTAN, Izhorskaya 13/19, Moscow, 127412 Russia
Fax: (095)485 7990, e-mail: kanel@icp.ac.ru

The experimental techniques and theoretical background plus some examples of


material strength measurements at sub-microsecond load durations are discussed.
The presented experimental data for engineering metals and alloys and for metal
single crystals demonstrate the effects of material grain structure, and temperature on
the resistance to high-rate rupture.

Keywords: dynamic tensile strength, shock waves, steels, aluminum, magnesium,


molybdenum single crystals, zinc single crystals.

Nomenclature

cb bulk sound velocity;


cl longitudinal sound velocity;
cF the velocity of spall signal front;
cp, cv heat capacities under constant pressure and constant volume, respectively;
p pressure;
t time;
T temperature;
Tm melting temperature;
u particle velocity;
ufs free surface velocity;
V specific volume;
Γ Gruineisen coefficient;
ρ density;
σx longitudinal stress component;
σ* spall strength;

1
Introduction

The dynamic tensile strength of materials at load durations of a few microseconds or less
is studied by analyzing spall phenomena under shock pulse loading. Spalling is the process of
internal rupture of a body due to tensile stresses generated as a result of a compression pulse
reflected from the free surface.
Numerous investigations of spallation have been performed for various practical
applications as well as for research into the physics of the strength of materials. The spallation
method has been applied to ductile and brittle materials, metals, inorganic single crystals and
glasses, elastomers and liquids over a wide range of load intensities and durations. Different
aspects of spall fracture investigations and their theoretical backgrounds are discussed in ref.
[1]. The principal content of investigations into spallation phenomena that have been
conducted are: (i) metallographic examinations of the spall zones in recovered samples in
order to obtain the correlation of the degree of damage with various features of stress history
[2,3,4] and (ii) instrumental measurements of the resistance of materials to dynamic fracture
[5,6]. In the former case the result is information about the ability of a material to withstand
shock loading, about the mechanism of a damage nucleation and development, and a
statistical description of these processes [4]. Dynamic measurements (free surface velocity
profiles or pressure at the interface with a soft material) during shock-wave loading give the
most appropriate and accurate data about the stresses present during the spall process.
During the dynamic fracture process, many microvoids or microcracks, more or less
simultaneously, undergo nucleation, growth, and coalescence in a volume of material to form
a failed or spalled region. Experiments with plane shock waves provide an unique base of
information about the strength of solids under small one-dimensional strain and stress states
close to three-dimensional tension. Neither the surface of the body nor isolated coarse defects
contribute to the main development of the spall fracture. One can say that shock-wave testing
permits one to measure the fundamental strength properties of matter.
This paper is devoted to discussing the methodology and capabilities of the technique to
measure spall strength, and the factors governing the high-rate fracture of metals and alloys
under such conditions. Some experimental data obtained earlier and their interpretation are
discussed with the main objective of outlining a possible place for shock-wave tests in the
strength of materials field.

Methodology of spall strength measurements.

Instrumental measurements of resistance to dynamic fracture, or the so-called spall


strength techniques, are based on the analysis of wave profiles recorded under conditions of
the loading of a sample by a plane shock wave. The shock-wave pulses of varied peak stresses
and durations are created in plane samples by a high velocity impact of a flyer plate, or by
detonation of a high explosive charge, or by pulsed laser or particle beams. Modern diagnostic
techniques permit one to record the wave profiles, particularly the free surface velocity
histories, with a high accuracy and sub-nanosecond time resolution.
Figure 1 shows examples of measurements for the high strength Fe-Cr-Ni-Mo steel
35Kh3NM [7]. The free surface velocity profile practically replicates the form of the
compression pulse in the sample if the load intensity is small (profile 1). The elastic-plastic
compression wave and the following complete unloading are recorded. A small velocity
hysteresis is explained by the hysteresis of cycle of elastic-plastic deformation.
Tensile stresses developed in the body after reflection of the compression pulse by the
free surface increase with an increase of shock intensity. When the peak tensile stress reaches

2
a critical magnitude, the nucleation and
growth of fracture are initiated. As fracture 500
develops, the tensile stresses relax to zero. 3
400

Free Surface Velocity, m/s


As a result, a compressive disturbance
called a "spall pulse" is produced in the free 5
300
surface velocity profile. Thereafter wave 4
reverberation is observed within the scab 2
200
between the free surface and the damage
zone. The period of velocity oscillation is a 100
measure of the thickness of the scab (or 1
spall plate). The velocity pullback from the 0
peak value to the value right ahead of the 0 1 2 3 4 5 6
Time, µs
spall pulse, ∆ufs, is a measure of the
incipient fracture strength of the material. Fig. 1. The free surface velocity profiles of a
Experiments show that increasing the shock high-strength structural steel samples impacted by
amplitude does not influence the magnitude flyer plates (curves 1, 2, and 4) or shock loaded
of ∆ufs. by detonation of explosive charges (curves 3 and
The peak free surface velocity, u0, and 5). Solid lines show free-surface velocity profiles
the free surface velocity just before the under loading in the rolling direction, dashed lines
are from experiments involving loading in the
arrival of the spall pulse, um, are determined lateral direction.
directly from the free surface velocity
profile. The tensile stress value just before spalling, σ∗, is then determined from an analysis
invoking the method of characteristics. From the acoustic approach, the following linear
approximation [5]
1
σ * = ρ o co ∆u fs , (1)
2
is used, where ∆u fs = uo − um , c0 is the sound velocity and ρ0 is the initial density.
For most solids, the free surface velocity profiles exhibit elastic-plastic properties. The
dynamic tensile strength of metals is usually much higher than the yield strength. We have to
decide which sound velocity should be used in Eq. (1) to calculate the tensile stress at a spall
plane. In the case of a one-dimensional process, weak perturbations propagate with a
longitudinal sound velocity, cl, if the deformation is elastic, and with the bulk sound velocity
cb < cl in the plastic deformation region.
Figure 2 shows an axial stress-particle velocity (σx-u) diagram for wave interactions
when a plane compressive pulse is reflected off the free surface of an elastic-plastic body.
The process of uniaxial compression is elastic until the stress reaches the Hugoniot elastic
limit or HEL. The slope of the initial elastic part of the Hugoniot below HEL in these
coordinates is dσ x / du = ρcl . The slope in the plastic deformation region above HEL is equal
to ρc b . Unloading from a shock-compressed state is initially elastic in both the incident and
reflected waves. Thereafter all expansion processes occur in the plastic region.
In order to determine the value of fracture stress we should look for the state
corresponding to the intersection point, K, of the terminal C- of the reflected rarefaction wave
and the C+ characteristic which create the minimum value in the measured free surface
velocity history. Since the spall pulse is a compression wave which propagates through the
extended material, the spall pulse front should propagate with a longitudinal elastic wave
speed cl whereas the incident rarefaction plastic wave ahead of it propagates with the bulk
sound velocity cb. Since the spall pulse front is considered as an elastic wave, it was
concluded [8,9] that in σx - u plane the C+ characteristic should correspond to the trajectory

3
Km′ in Fig. 2 which passes through the point σx = 0, u = um′ and has a slope dσx /du = -ρcl
while the trajectory corresponding to states along the terminal C- characteristic has a slope
ρcb. With this approach, the relationship by which to calculate the value of fracture stress σ*
is
1
σ * = ρ 0 cl ∆u fs . (2)
1 + cl c b
This relationship does not account for the spall plate thickness. However, the distortion of the
wave profile obviously should increase with increasing distance as a result of different
propagation velocities. The interaction of the elastic compression wave with the plastic
rarefaction wave, both of which propagate in the same direction, requires a more detail
analysis.

σx Pressure

Elastic release
in the reflected wave Compression

F
c
Incident

t=
dh F
cl

/d
unloading Plastic release =

Time
HEL /dt
dh
m m' O'
0 Elastic up
compression
after spall
-+
Plastic
Rarefaction

b
rarefaction

=c
/dt
dh
K

Distance

Fig. 2. States of an elastic-plastic material at reflection of compression pulse by the free


surface.
Fig. 3. C+ characteristics of a plastic rarefaction wave followed by an elastic compression in
the distance, h, - time, t, coordinates. F is a trajectory of the compression wave front; cl and cb
are the longitudinal sound speed and the bulk sound speed, respectively; signs “+” and “-“
denote the flow areas ahead of the compression wave front and behind it.

The time-distance diagram presented in Fig. 3 shows C+ characteristics describing a part


of the incident plastic rarefaction wave followed by an elastic recompression wave. The
recompression wave front propagates with a velocity, cF, which is within the range cl ≥ cF ≥
cb. Accounting for the conservation of momentum, the stress gradients along the trajectory F
are
dσ x dσ x
= σ& x + − c F ρu& + and = σ& x − − c F ρu& − ,
dt F dt F

where the indexes “+” and “−” denote parameters on the right and left sides of the trajectory
F, respectively; f& = ∂f ∂t . Accounting for the mass conservation equation, the particle
velocity gradients along the trajectory are
du c F σ& x + du c F σ& x −
= u& + − 2 and = u& − −
dt F ρcb dt F ρcl2

4
where σ& x + = −V&+ ⋅ ρ 2 cb2 , σ& x − = −V&− ⋅ ρ 2 сl2 , V is the specific volume. Since the stress and the
particle velocity derivatives along the trajectory F should be the same on both sides of F, we
are coming to a set of two equations for the front propagation velocity, cF, and the stress and
the particle velocity alteration rates just ahead of the recompression front and behind it.
Solving this set, we receive the propagation velocity for the recompression wave front:

σ& x + − σ& x −
c F = cb cl (3)
σ& x + cl2 − σ& x − cb2 −
which should be true for σ& x + and σ& x − of different signs. According to this solution, the
recompression wave front propagates with the velocity of the longitudinal elastic wave, cl,
when the stress gradient ahead of it, σ& x + , is zero or when the recompression wave is a shock
discontinuity ( σ& x − → ∞ ).
When a spall fracture occurs as a result of the reflection of a triangular compression
pulse, interference of the incident and the reflected rarefaction waves supports constant tensile
stress values in each cross-section of the body until the spall pulse arrives. As a result, the
spall pulse front propagates with the velocity cF = cl independent of its gradient. In this case
all characteristics of the spall pulse reach the rear free surface and the spall strength may be
calculated equivalently using both the approach given by Eq. (2) or the approach based on
consideration of states just ahead of the spall pulse which gives the relationship
1 h h 
σ * = ρcb (∆u fs + δ ) , δ =  sp − sp  ⋅ u&1 . (4)
2  cb c F 
Where u&1 is the free surface velocity derivative ahead of the spall pulse, cF = cl for the load
pulse of triangular shape an hsp is the spall plate thickness.
The shock load pulses of approximately triangular shape are created in solids, for
example, by detonation of an explosive charge. A more common way of shock loading is by
the impact of a plate which is associated with compression pulses of approximately
trapezoidal shape. In this case the stresses in cross-sections ahead of the spall pulse front are
not constant, so c F < cl . This means that some leading characteristics of the spall pulse will
disappear with time as is shown in Fig. 3. As a result, Eq. (2) is not valid even for a shock-like
spall pulse, so only Eq. (4) may be used. To find the δ correction value, we should calculate
cF accounting for σ& x + ≈ ρcb u&1 / 2 , σ& x − = ρ cl u& 2 / 2 near the spall plane and σ& x + ≈ 0 near the
rear free surface of the sample plate. Here u& 2 is the free surface velocity gradient in the spall
pulse front.
The δ correction value is smallest for the case of a triangular shape of load pulse profiles.
Inaccuracy in the δ value caused by extrapolations of u&1 and u& 2 is also the smallest in this
case, so it is recommended to carry out spall strength measurements using shock load pulses
of a triangular shape.

Spall strength of engineering metals and alloys

Metals and alloys are the materials most thoroughly investigated with regard to spall
phenomena. Table 1 summarizes known data on the dynamic yield strength and the spall
strength of some selected engineering materials. Let us consider several examples which
demonstrate the main peculiarities of their spall fracture.

5
Table 1. Yield strength and spall strength of some metals and alloys.
Dynamic yield Spall strength,
HEL, GPa
Material strength, Yd, GPa Refs.
GPa
Aluminum 2024 0.6 0.29 1.15 [20]
Aluminum AMg6M 0.38 0.18 0.8 to 1.1 [21]
Aluminum AD1 1.1 to 1.3 [11]
Magnesium Mg95 0.8 to 1.0 [11]
Magnesium Ma1 0.2 0.1 0.8 to 0.9 [21]
Copper M2 0.8 to 1.6 [21]
Copper OFHC 0.12 1.2 [22]
Tungsten 4.0 2.4 0.7 [20]
Titanium Ti-6Al-4V 2.0 0.92 3.6-4.4 [21]
Tantalum 0.77 4.4 [22]
Armco iron 0.95 0.56 1.4-1.7 [21]
As received 40Kh chromium-doped [10]
structural steel, 17 to 19 HRC 1.7 0.83 2.3
Quenched 40Kh steel, 45 to 54 HRC [10]
2.15 0.92 4.2

The free surface velocity profiles for steel samples of different orientation, which are
presented in Fig. 1, demonstrate the influence of load direction on the resistance to spall
fracture. The wave profiles shown as solid lines were obtained under loading in the rolling
direction, while profiles shown as dashed lines were obtained under loading in the lateral
direction. The resistance to dynamic fracture is 4.0 to 4.4 GPa in the case of loading in the
rolling direction and by 0.5 to 0.6 GPa (that is about half of the yield strength) less in the case
of loading in the lateral direction. More fast decay of the velocity oscillations in the latter case
correlates with a more highly developed fracture surface.
One effect of the material texture orientation is fracture nucleation at inclusions and other
relatively coarse defects which are concentrated on grain boundaries. To some degree, the
material microstructure can be changed by the plastic deformation preceding fracture and by
polymorphous transformations. It is known that reversible α to ε (b.c.c. to h.c.p.)
polymorphous transformations occur in iron and steels under compression at 13 GPa pressure.
Shock-wave experiments with as-received and quenched chromium-doped structural 40Kh
steel have been carried out [10] at peak stresses below and above the transformation pressure.
The measurements have shown that at lower peak stresses the strengthening effect of heat
treatment manifests itself more on the dynamic tensile strength than on the dynamic yield
strength. For both states of this steel the dynamic tensile strength increases as a result of the
α-ε polymorphous transformation. When the peak stress much exceeds the transition pressure,
the spall strength of as-received steel approaches the strength of quenched samples. It is
known that the reversible phase transformation causes a reduction in grain size with the
formation of a martensite structure, and, as a result, provides a sharp increase in hardness. It
seems, that these effects hinder the growth of microcracks, which can be the reason for the
increase in spall strength. Quenched samples initially have a fine grain martensite structure,
therefore the strengthening effect due to the reversible phase transformation is lower in these
samples.
Beyond micro-structural transitions, the intensity of the shock compression preceding to
tension does not produce any notable effect on the resistance to spall fracture. Figure 4 shows

6
0,8

1,4 Aluminum AD1 4,8


Sn
0,6

Spall Strength, GPa


Spall Strength, GPa

17,4
1,2 10,4
34,5
5 0,4
50

1,0 5,4
0,2

0,8 0,0
104 105 . 106 0 10 20 30 40
-1
Strain Rate V/V0, c Peak Shock Stress, GPa

Fig. 4. The spall strength of aluminum AD1 as a function of the pressure and the
decompression strain rate. The peak shock stresses are indicated (in GPa) by numbers at the
points.
Fig. 5. Results of measurements of the spall strength of tin [13, 14].

experimental data on the spall strength of aluminum AD1 (analogous to Al 1100) over a wide
range of peak pressures and strain rates V& V0 = − u&1 2cb [11]. The spall strength, in fact,
does not depend on the peak shock pressure in the range up to 35 GPa. Two experiments at a
peak pressure of 50 GPa were not very accurate. The same conclusion follows from
experiments with aluminum and titanium alloys over a wide range of peak stresses [12]. Since
the spall strength is sensitive to the material grain structure and texture and is not sensitive to
the preceding strain from the shock compression wave, the damage nucleation sites in
engineering materials are relatively coarse structural defects such as inclusions, grain
boundaries, etc. Dislocations and other microscopical defects produced by the high-rate
deformation obviously are not active.
On the other hand, the shock compression is accompanied by irreversible heating of the
material. Figure 5 shows the spall strength data [13, 14] for tin as a function of the peak shock
stress. The data show a substantial drop in the spall strength when the peak shock pressures
approach values sufficient to achieve melting after the shock-wave compression and
unloading process.
The damage rate is approximately equal to the product of the concentration of damage
nucleation sites times their average growth rate and cannot be arbitrarily large. As a
consequence, the spall strength is not a material constant but depends on the load duration or
the strain rate, as can be seen in Fig. 4. As a result of significant overstressing of the material,
numerous small defects are initiated when the load is applied at a high rate. It can be pointed
out in this connection that irregularities observed on the fracture surface become finer with a
decrease in the load pulse duration. The rate of damage that can be observed in the wave
profiles is correlated with the decompression rate in the incident load pulse. Actually, the
spall pulse can be seen on the free-surface velocity profile if the fracture is fast enough. A
slower fracture process leads to an increased rise time of the spall-pulse front. This means that
not only the fracturing stresses but also the corresponding fracture rates can be evaluated from
the test measurements.

7
4. Spall strength of metal single crystals

Experiments with single crystals provide information about the conditions for fracture
nucleation from dislocations and point defects. Such investigations were carried out with
copper [15], molybdenum [16], and zinc [17]. Figures 6 presents the spall strength data for
molybdenum single crystals in comparison with the results of similar measurements for
polycrystalline molybdenum and previously deformed molybdenum single crystals. The spall
strength of molybdenum single crystals is twice and that of copper single crystals is thrice of
polycrystalline samples. Thus, in the case of polycrystalline materials, microdefects may not
serve as fracture nucleation sites because the stress relaxation produced by damage growth at
coarse defects does not permit the required level of tension.
Plastic deformation of these single crystals during shock compression and unloading
preceding tension produced numerous dislocations and points defects which could reduce the
resistance to tensile fracture. In the case of zinc it was possible to compare the dynamic
tensile strength of deformed and undeformed single crystals. Zinc has an hexagonal close-
packed crystal lattice with a large degree of anisotropy. The main slip plane for h.c.p. crystals
_
is the basal plane (0001); dislocation slip and twinning in the pyramidal ( 112 2 ) plane requires
a much higher shear stress. In experiments, the collision planes were the basal plane (0001)
_
and the prismatic plane ( 10 1 0 ) which is perpendicular to the basal plane. At the orientations
chosen, there was no shear stress in the basal plane, so only secondary slip systems were
activated.
Fig. 7 shows the typical free surface velocity histories recorded at the shock compression
of zinc single crystals of different orientations. The bulk sound velocity in zinc is cb = 3.03
km/s; the longitudinal sound velocity in the direction perpendicular to the hexagonal axis of
the crystal is cl = 4.73 km/s while along the axis direction it is very close to the bulk sound
_
velocity. As a result, for the samples impacted on the ( 1010 ) plane, which is perpendicular to
the hexagonal c-axis, the shock waves were split with the forming of the elastic precursors
and the plastic compression wave, whereas for the axial loading direction purely elastic shock
waves of large peak stresses were recorded. The rise-time in the elastic shock wave and the
elastic precursor front is close to the resolution limit of 2 to 4 nanoseconds. The rise-time in
the plastic shock front of 6 GPa peak stress is around 80 to 90 nanoseconds.
It is known that under normal conditions the zinc single crystals can be readily cleaved
along the basal plane. The shock-wave profiles shown in Fig. 7, however, demonstrate a high

20 600
Spall Strength, GPa

Single (1010), 1.7mm,


Free Surface Velocity, m/s

10 500 (0001), 1.55 mm Impactor 0.85 mm


crystals Deformed
single
400
5 crystals
4
300
3 _
(1010), 0.5 mm
2 200
Polycrystals
1 Mo 100

0.5 0
0,0 0,2 0,4 0,6
104 105 106 107 108
. Time, µs
V/V0, c-1

Fig. 6. The dependence of molybdenum spall strength upon the strain rate in unloading part
of initial shock pulse.
Fig. 7. Free surface velocity profiles for zinc single crystal samples. Impactor and sample
thicknesses are mentioned.

8
1,5
p

Spall Strength, GPa


1,0 Mg Al S

M
0,5

- Aluminum AD1
V(Tm)
- Magnesium Mg95
0,0 0 V(T0) V
0,4 0,6 0,8 1,0
T/Tm
K

Fig. 8. The spall strength of aluminum and magnesium as a function of the temperature
divided by the melting temperature. Dashed lines correspond to Eq. (9) for aluminum and
magnesium.
Fig. 9. The scheme of mutual positions of the isentrope, S, and the melting curve, M, in the
pressure-specific volume plane.

spall strength for dynamic fracture on the (0001) basal plane: the spall strength measured at
shock loading in the axial and the transverse directions are 2.1 ± 0.2 GPa and 1.33 ± 0.3 GPa,
respectively. Thus, we find the unexpected result that the spall strength is larger for fracture
on the weakest plane. It seems that the main reason of high strength along (0001) plane is that
there was no plastic deformation under shock compression so the fracture was initiated in a
perfect initial structure. Experiments with molybdenum [16] also show some decrease in the
spall strength of pre-strained single crystals as compared to the undeformed samples.

Athermicity of the dynamic strength and the melting threshold.

It is well-known that under normal conditions both the yield strength and the tensile
strength are strong functions of the temperature. For low rates of mechanical loading, the
dislocation motion is aided by thermal fluctuations [18]. A transition to athermal plasticity
occurs at high strain rates, when the applied stress is high enough to overcome the usual
dislocation barriers without any aid from thermal fluctuations. Since the fracture process
involves the plastic flow around growing voids, a transition to athermal fracture should be
expected at similar strain rates.
Figure 8 shows the spall strength dependence on the initial temperature for aluminum
AD1 and magnesium Mg95 measured at a peak compressive shock pressure of 5.8 GPa and
3.7 GPa, respectively [11]. In the figure, the normalized (i.e. homologous) temperature is
T/Tm, where Tm is the melting temperature in Kelvins. In general, the spall strength maintains
almost a constant value with increasing initial temperature up to 85 to 90% of the melting
temperature. After that, a precipitous drop in the spall strength occurs as temperature
approaches the melting point both for aluminum and magnesium. Since at the same time the
dynamic yield strength does not decrease near the melting point, it is natural to associate the
precipitous decrease in spall strength with melting under expansion into the negative pressure
region.
In the framework of a common approach the rarefaction of shock-compression matter
may be considered as isentropic. Figure 9 illustrates the mutual position of the isentrope and

9
the melting curve in pressure, p,
2,5 volume, V, coordinates. Since the
melting curve is steeper than the
2,0 isentrope, the rarefaction shifts the
state of solid matter toward the
Spall Strength, GPa

melting curve. Intersection of these


1,5
two curves means the beginning of
Zn single crystals melting at the temperature T < Tm0. In
1,0
Impact plane (0001) as much the bulk strength of liquids is
Tm much less than that of solids, the
0,5 strength should drop practically to
zero when the material begins to melt.
0,0 Let us estimate in a linear
0 100 200 300 400
approach the pressure at intersection
Temperature, oC point K in Fig.9. The melting
Fig. 10. Comparison of the estimations of boundary line in these coordinates has
strength using the Eq. (9) (dashed line) with the a slope
experimental spall data for zinc single crystals.
dV  ∂V  dTm  ∂V 
=  +  (5)
dp  ∂T  p dp  ∂p  T
where the melting temperature as a function of pressure, Tm(p), is known. The linearized
isentrope of solids with a temperature T0 at p = 0 is
 ∂V   ∂V 
V = V (Tm 0 ) + (T0 − Tm 0 ) ⋅   +   p , (6)
 ∂T  p =0  ∂p  S
where Tm0 is the melting temperature at zero pressure. Taking into account the relationships
between the isothermal compressibility and isentropic compressibility
 ∂V  c  ∂V   ∂p   ∂V 
  = V ⋅   and c p − cV = −T     , (7)
 ∂p S c p  ∂p T  ∂T V  ∂T  p

where cV and cp are heat capacities under constant volume and constant pressure,
respectively, we obtain the condition at the point of intersection of the isentrope and the
melting curve:
(T0 − Tm 0 )
p= , (8)
dTm 2
− ΓT ρc s
dp
where Γ is the Gruineisen coefficient, and cs is the sound velocity. The temperature T in Eq.
(8) can be replaced by the pressure that finally gives the equation for the pressure at the
intersection point
dTm  dT 
ρc s2 − Γ ⋅  Tm 0 + p m  − (T0 − Tm 0 ) ⋅ ρc s2 p = 0 (9)
dp  dp 
Figures 8 and 10 compare the estimations of strength using Eq. (9) with the experimental
spall data for polycrystalline Al and Mg and zinc single crystals. The thermophysical
properties of these metals used in the calculations are listed in Table 2. To compare these data
we should account for a temperature growth of ~10° due to yielding and viscosity effects in
the compression phase preceding tension.

10
Table 2. Thermophysical properties of metals [23-26]

ρ, g/cm3 dTm/dp, cs, km/s dσ ∗


Material Tm, K Γ
(T ≈Tm) grad/GPa (T ≈Tm) dT0
MPa/grad
Aluminum 933.1 2.56 64.1 4.6 2.15 -32.1
Magnesium 922.5 1.642 75.0 4.7 1.57 -25.6
Zinc 692.5 6.81 48.1 3.16 2.24 -35.4

The estimations roughly agree with the experimental data and deviate from them both to
the higher (Al, Mg) and lower (Zn) strength values. The disagreement may be associated with
inaccuracies in the mechanical and thermo-physical data used, as well as with the nature of
the shock-wave loading. In the case of polycrystalline metals, hot spots may be formed as a
result of partial localization of energy of shock wave on microvoids and inclusions.
Obviously, the shock-wave heating of single crystals is much more homogeneous, however it
is not quite clear what is the contribution of anisotropy into the overall response.
The observed melting threshold opens a way to investigate the melting and other phase
transitions in the negative pressure region. In this context, one should mention that an ultimate
tensile strength is realized when single-crystalline samples are shocked in the nanosecond
load duration range. On the other hand, such results probably indicate the physical reason for
the ultimate strength of solids. In this respect we may note the results of large-scale molecular
dynamic simulations of fracture nucleation in copper single crystals [19]. It has been shown
that voids form in an ideal crystal as a result of the lattice becoming unstable in small regions
which can be interpreted in terms of homogeneous nucleation of melted spots.

Conclusions

Spall strength measurement techniques have been developed initially with the main aim
of characterizing those materials used for military applications. We hope such techniques can
be useful in the fields of engineering and material science and with respect to the physics of
material strength. The techniques and the theoretical background of the method are well
advanced. In this paper we have presented several examples which demonstrate the effects of
material grain structure, preceding plastic deformation, and temperature on the resistance to
high-rate rupture. The most important thing now is to define how to relate these sub-
microsecond measurements with the material strength properties for large load durations and
for cycled loads which are of more interest to engineers.

References

1. L. Davison, D.E. Grady, and M. Shahinpoor (1996). High Pressure Shock Compression of
Solids - II. Dynamic Fracture and Fragmentation. Springer-Verlag New York, Inc.
2. J.H. Smith (1962). Three low-pressure spall thresholds in copper. - ASTM Spec. Techn.
Publ., No 336, 264-281.
3. M.A. Meyers and C.T. Aimone (1983). Dynamic fracture (spalling) of metals. - Progress
in Material Science, 28, 1-96.
4. D.R. Curran, L. Seaman, and D.A. Shockey (1987). Dynamic failure of solids. - Phys.

11
Reports, 147(5&6), 253-388.
5. S.A. Novikov, I.I. Divnov, and A.G. Ivanov (1966). The study of fracture of steel,
aluminum and copper under explosive loading. - Phys. of Metals and Metallography
(USSR), 21(4), 608-615.
6. L. Davison and R.A. Graham (1979). Shock compression of solids. - Phys. Reports, 55(4),
255-380.
7. V.D. Glusman, G.I. Kanel, V.F. Loskutov, V.E. Fortov, and I.E. Khorev (1985). The
resistance to deformation and fracture of steel 35Kh3NM under shock loading conditions -
Problems of Strength (USSR), No 8, pp.52-57.
8. G.V. Stepanov (1976). Spall fracture of metals by elastic-plastic loading waves. Problems
of Strength (USSR), No 8, 66-70.
9. V.I. Romanchenko and G.V. Stepanov (1980). Dependency of critical stresses on the load
time parameters at spalling in copper, aluminum, and steel. J. Appl. Mech. Tech. Phys.,
21(4), p.141.
10. S.V. Razorenov, A.A. Bogach, and G.I. Kanel (1997). The effect of heat treatment and
polymorphic transformations on the dynamic strength of steel 40 Kh. Phys. Of Metals and
Metallography, 83(1), 100-103.
11. G.I. Kanel, S.V. Razorenov, A.A. Bogatch, and A.V.Utkin, V.E.Fortov, and D.E.Grady
(1996). Spall Fracture Properties of Aluminum and Magnesium at High Temperatures. J.
Appl. Phys., 79(11), 8310-8317.
12. G.I. Kanel, S.V. Razorenov, and V.E. Fortov (1987). Spall strength of metals over a wide
range of amplitude of the shock load. Dokl. AN SSSR, 294(2), 350-352.
13. Grady, D.E. (1988). The spall strength of condensed matter. - J. Mech. Phys. Solids,
36(3), pp.353-384.
14. G.I.Kanel, S.V.Razorenov, A.V.Utkin, and D.E.Grady (1996). The Spall Strength of
Metals at Elevated Temperatures. in: Shock Compression of Condensed Matter - 1995.
Eds.S.C.Schmidt, W.C.Tao. AIP Conference Proceedings 370, 503-506.
15. S.V. Razorenov and G.I. Kanel (1992). The strength of copper single crystals and the
factors governing metal fracture in uniaxial dynamic stretching. Phys. Of Metals and
Metallography, 74(5), 526-530.
16. G.I. Kanel, S.V. Razorenov, A.V. Utkin, V.E. Fortov, K. Baumung, H.U. Karow, D.
Rush, and V. Licht (1993). Spall Strength of Molybdenum Single Crystals. J. Appl. Phys.,
74(12), 7162-7165.
17. A.A. Bogatch, G.I. Kanel, S.V. Razorenov, A.V.Utkin, S.G. Protasova, and V.G. Sursaeva
(1998). The resistance to shock-wave deformation and fracture of zinc single crystals at
elevated temperatures. Sov. Phys. - Solid State, 40(10), 1849-1854.
18. A. Kumar and R.G. Kumble (1969). Viscous drag on dislocations at high strain rates in
copper. J. Appl. Phys. 40(9), 3475-3480.
19. J. Belak (1998). Molecular Dynamics Simulation of High Strain-Rate Void Nucleation
and Growth in Copper. In: Shock Compression of Condensed Matter - 1997, Edited by
S.C. Schmidt, D.D. Dandekar, J.W. Forbes, AIP Conference Proceedings, 429, 211-214.
20. C.E. Morris (Ed.) (1982). Los Alamos Shock Wave Profile Data. University of California
Press.
21. G.I. Kanel, S.V. Razorenov, A.V. Utkin, and V.E. Fortov (1996). Shock-Wave
Phenomena in Condensed medias. “Yanus-K”, Moscow. (in Russian).
22. Cochran, S., and D. Banner (1977). Spall studies in uranium. J. Appl. Phys., 48(7), 2729.
23. M.W. Guinan and D.J. Steinberg (1974), Pressure and Temperature Derivatives of the
Isotropic Polycrystalline Shear Modulus for 65 Elements. J. Phys. Chem. Solids, 35, 1501-
1512.
24. T. Gorecki (1979). Vacancies and Generalized Melting Curves of Metals. High

12
Temperatures - High Pressures, 11, 683-692.
25. I. N. Frantcevich, F.F. Voronov, S.A. Bacuta (1982). Elastic constants and elastic modulus
of metals and non-metals (Reference book). Kiev, Naukova dumka.
26. G. Simmons (1965). Single Crystal Elastic Constants and Calculated Aggregate
Properties. J. Grad. Res. Cent., 34(1-2), 1-269.

13

You might also like