The Failure and Fracture Modes of 35CrMnSi Steel Projecti - 2019 - Engineering F

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Engineering Failure Analysis 97 (2019) 617–634

Contents lists available at ScienceDirect

Engineering Failure Analysis


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

The failure and fracture modes of 35CrMnSi steel projectile with


T
simulated charge penetrating rock at about 1000 m/s

Qiran Suna,b, Yuxin Suna, , Ruiyu Lic, Guoqiang Dengd, Jinsheng Hud
a
National Key Laboratory of Transient Physics, NJUST, Nanjing 210094, Jiangsu, China
b
Department of Mechanical Engineering, Southern Methodist University, Dallas 75275, TX, USA
c
China Ship Development and Design Center, Wuhan 430064, Hubei, China
d
No. 4 Engineering Research Institute, General Staff Headquarters of PLA, Beijing 100850, China

A R T IC LE I N F O ABS TRA CT

Keywords: To reveal the projectile failure and fracture modes during high-velocity impact, projectiles made
Projectile fracture mechanism of 35CrMnSi with the same shape but different shell thickness are designed and then the ex-
High-velocity penetration periments of penetrating high-strength rock target at about 1000 m/s are carried out. The ex-
Petal shaped fragments perimental results indicate that all projectiles are completely fractured with petal shaped frag-
Numerical analysis
ments produced and fail to effectively penetrate the rock target while only the target surface is
Mott failure distribution
comminuted. Besides the discussion on projectile material strength failure based on experimental
results, projectile structure damage in the form of dynamic plastic buckling failure is also applied
and gives a satisfactory explanation for the fracture pattern of petal shaped fragments with three-
hinges structures employed. In addition, by taking the failure model of Mott stochastic dis-
tribution into account for the projectiles, numerical simulations are carried out to analyze the
projectile fracture modes by Autodyn-3D with different discrete forms and exhibit high cred-
ibility. Furthermore, the influences of the filling and different impact conditions including ve-
locities, angles of attack (AOA) and incidence angles on the projectile fracture modes are dis-
cussed. In general, the experimental results and the adopted numerical method with high
credibility are able to reveal the effect of projectile structure damage on fracture modes and can
be referred to further study on the projectile fracture during high-velocity penetration.

1. Introduction

A lot of efforts have been made to analyze various target material failure mechanism subjected to impacts [1–3]. Especially, for
low-velocity projectile penetration (< 1000 m/s) into geological materials, such as soil [4–7], concrete [4,5,8–10] and rock [4,5,11],
by assuming the projectile as a rigid body, the target failure mechanism and projectile penetration ability can be effectively in-
vestigated. However, with the increase of the projectile impact velocity and the target strength, the stress field with larger strains and
higher strain rates will form, which can cause the obvious phenomena, such as bending failure or fracture of the projectile [12–19].
Therefore, Darrigade et al. [12] concluded that the intensive load may lead to severely structural destruction when a slender thin-
wall projectile impacting into steel plate or high-strength concrete at high speed. Also for the penetration experiments into limestone
targets (uniaxial compressive strength fc=60 MPa) conducted by Frew et al. [13], the projectile bending failure was observed in the


Corresponding author at: National Key Laboratory of Transient Physics, Nanjing University of Science and Technology, No. 200 Xiaolingwei
Street, Nanjing 210094, China.
E-mail address: yxsun01@163.com (Y. Sun).

https://doi.org/10.1016/j.engfailanal.2019.01.057
Received 2 October 2018; Received in revised form 10 January 2019; Accepted 10 January 2019
Available online 11 January 2019
1350-6307/ © 2019 Elsevier Ltd. All rights reserved.
Q. Sun et al. Engineering Failure Analysis 97 (2019) 617–634

Fig. 1. Schematic of projectile geometry.

cases of impact velocity vi≥1649 m/s. While for those into granite targets (fc=154 MPa) carried out by Zhang et al. [14], all pro-
jectiles performed obvious bending failure with vi ranging from 332 to 988 m/s and even fractured when vi > 797 m/s. Recently,
Kong et al. [15], Hu et al. [16] and Chen et al. [17] conduct much more discussion on projectile deformation and erosion during the
penetration into concrete targets. For metal targets, Rakvåg et al. [18,19], Xiao et al. [20], Chen et at. [21] and Wang et al. [22]
studied the deformation and fracture modes of steel projectiles during impact experimentally and numerically. It is noteworthy that
the analyses on the structure failure, conducted by Jones [23], are implemented by Darrigade et al. [12] and Chen et al. [21] to
explain the projectile fracture mechanism during the penetration. In fact, due to the bending failure or fracture of the projectile, its
penetration ability is seriously limited. Considering the investigations focused on projectile deformation and fracture during pene-
tration are less widely than those focused on rigid projectile penetration, it is necessary to reveal the projectile fracture mechanism
during penetration and then contribute to the improvement of the penetration ability.
In order to reveal the projectile fracture mechanism during high-velocity penetration into rock, projectiles with the same shape
but different shell thickness are designed and then the experiments with vi≈1000 m/s are carried out. Based on the experimental
results, the fracture pattern of petal shaped fragments are theoretically analyzed. In addition, the corresponding numerical model of
penetrating rock target is established with coupled SPH-FEM by Autodyn-3D. By taking the failure model of Mott stochastic dis-
tribution into account for the projectiles, the numerical simulations are performed and exhibit high credibility. Furthermore, the
influences of the filling simulated as charge and different impact conditions on the projectile fracture pattern are discussed.

2. Field tests

2.1. Description of projectiles and target for tests

For field penetration tests, projectiles including the shell, the filling and the back cover are designed, indicated in Fig. 1. The shell
as the bulk of projectile, made by 35CrMnSi steel alloy after heat treatment with the ultimate tensile strength reaching 1.5GPa
[22,24], is utilized herein for tests. In addition, the main chemical and mineral compositions of 35CrMnSi steel alloy are given in
Table 1. The back cover materials is 7039 aluminum alloy with the ultimate tensile strength of 0.335GPa. The filling material is
sulfur, employed to simulate the charge inside the projectile.
According to the variation of shell thickness, the whole shell can be divided into four part as follows: the ogive-nose part, the
transition part, the cylindrical part and the tail part, marked as a, b, c and d in Fig. 1. Four sub-caliber projectiles, having the same
shape but different shell thickness of the cylindrical part hc = 17mm (1# and 2# in Fig. 2a) and 22 mm (3# and 4# in Fig. 2a) with
the corresponding mass 8.35 kg and 9.95 kg, were launched by 125 mm caliber smoothbore gun at about 1000 m/s (Fig. 2b) with a
distance of 50 m away from the rock surface.
The rock target is constructed by natural mountain excavation (Fig. 2b). The rock material is a kind of rhyolite having almost the
same composition with granite. Besides, four cylindrical rock samples with 40 mm diameter and 80/20 mm height are acquired by
drilling and precision work. In addition, physical indices and statics mechanics parameters of rhyolite are obtained by the corre-
sponding average value of compression and split tests (Fig. 3), shown in Table 2.

2.2. Results of penetration tests

Fig. 4 shows the typical impact surface of the rock target. It can be found that the rock falling off from the target surface after

Table 1
Chemical and mineral compositions (weight %) of 35CrMnSi steel alloy.
Fe C Si Mn Cr Ni P S Cu

≥96 0.32–0.39 1.10–1.40 0.80–1.10 1.10–1.40 ≤0.030 ≤0.025 ≤0.025 ≤0.025

618
Q. Sun et al. Engineering Failure Analysis 97 (2019) 617–634

Fig. 2. Preparation for field test.

Fig. 3. The rock mechanics test.

high-velocity impact is quite regular and then the target surface forms obvious divergent cracks. The depth of crater caused by impact
is 53.2 cm and the new free surface with divergent cracks is about 102 cm in height and 91 cm in width.
However, projectiles with different shell thickness are all fractured and fail to effectively penetrate into targets (showed in
Table 3). Projectile fragments are dispersed in the test field after impact and only a part of them can be collected, presented in Fig. 5

619
Q. Sun et al. Engineering Failure Analysis 97 (2019) 617–634

Table 2
Quasi-static parameters of rhyolite and granite material.
Parameters Rhyolite Barre Granite [25]

Density ρ/(kg/cm3) 2.66 2.66

Uniaxial compressive strength fc/(MPa) 178 167.1


Uniaxial tensile strength T/(MPa) 7.9 7.3

Fig. 4. Typical impact surface of rock target (#3).

Table 3
Test results of impact rock.
Number hc/mm Damage of projectile vi/(m/s)

#1 17 Fractured 958
#2 17 Fractured 1043
#3 22 Fractured 983
#4 22 Fractured 991

Fig. 5. Projectile fragments collected from test field (#3).

(according to the most likely place where it belongs). Meanwhile, none of fragments split from the projectile transition part identified
by obvious varying shell thickness is collected. The comparison of the fragments in different columns in Fig. 5 indicates that, the
closer to the projectile tip the fragment location, the smaller the fragment and the more serious the projectile damage. Fragments
with large axial size (column 3 and 4 in Fig. 5) are mainly split from the projectile rear part. Especially, the fragments in column 5 of
Fig. 5 can be pieced together to almostly form the whole projectile tail, which means the projectile rear part was divided into four (or
more) pieces during penetration. In addition, only a few fragments split from the impactor nose (column 1 in Fig. 5) confirm that the

620
Q. Sun et al. Engineering Failure Analysis 97 (2019) 617–634

Fig. 6. Typical eversion of fragments.

impactor nose is fractured into small fragments due to severe plastic deformation.
Furthermore, the fragments 5# and 8# with obvious eversion present as petal shape (Fig. 6), which indicates the projectile shell
has large radial displacement. Combining Fig. 6 with Fig. 7, the deduction that the projectile is subjected to unstable tensile failure
and typical shear failure can be confirmed accordingly by irregular circumferential fracture surface and the 45° angle on the annular
fracture surface.
It is noteworthy that the impactor nose breaks into small fragments completely and only a few of them can be collected, which is
significantly different from the impactor nose almost remaining intact in the experiment of penetrating the thick metal target in the
Ref. [21] (See Fig. 8). The main reason is, for the rock target with semi-infinite thickness, the impactor nose is resisted even eroded
during penetration, leading to the impactor nose crush and unable to maintain integrity. While the impactor nose penetrating thin
metal target has already completed the perforation before the existence of severe deformation due to the short penetration time.

3. Analysis on the projectile fracture mechanism

During high-velocity penetration, the material strength failure or the projectile structure damage, or both, may occur. Especial for
thin-wall projectile, local unstable deformation due to projectile bulking (Fig. 9) is easy to take place, which is regarded as structure
damage and investigated by Darrigade et al. [12] and Chen et al. [21]. Once local unstable deformation occur, the projectile will lose
structure integrity and then fracture [12,21]. Overall, both factors play different significant roles in the projectile penetration ability.
In addition, due to the variation of the projectile shell thickness and the incomplete collection of fragments in test field, both damage
modes should be studied and then the corresponding effect will be comprehensively and accurately distinguished. Considering the
influence of the material strength failure on the projectile fracture pattern has been confirmed by the observation of severe plastic
deformation in the collected fragments, only further analysis on the other factor of structure damage is needed.

3.1. Material strength failure analysis

Experimental results (Fig. 7) demonstrate that the projectiles have complex shearing or tensile failure during the impact. In order
to further explore and verify the material failure mechanism of the projectile, scanning electron microscopy (SEM) analysis of
specimens cutting form the axial fracture surface (Fig. 6b) and the front fracture surface (Fig. 7a) in a typical fragment 8# is
performed.

Fig. 7. Typical circumferential fracture sections of fragment 8#.

621
Q. Sun et al. Engineering Failure Analysis 97 (2019) 617–634

Fig. 8. Experimental results with intact impactor nose [21].

Fig. 9. Schematic of projectile buckling during penetration [12].

Fig. 10 is the SEM image of specimens cut from fragment 8# with the axial fracture surface (corresponding to Fig. 6b). There are
many obvious “band-like” layers in Fig. 10, each of which has a smooth flat surface and perform obvious directivity [26]. It indicates
the fracture occurs along a certain crystal plane, and then is flattened by strong pressure to form a shear band [26]. This is a typical
feature for high-strength material with shear fracture.
Fig. 11 is the SEM image of specimens cut from fragment 8# with the circumferential fracture surface (corresponding to Fig. 7a),
it shows the fracture of crystal grains. Large cavities are also produced due to the engagement of intercrystalline microscopic voids
caused by the grain boundary slip. However, the topographical representation image is not totally the same as tensile dimple, which
is caused by the deformation dominated by dislocation motion at low temperature. It indicates that the destruction on the cir-
cumferential fracture surface of fragment 8# is caused by the grain boundary slip under high temperature and high strain. It is
embodied macroscopically as a tensile failure feature at high temperatures.
In general, the SEM images of fracture surface exhibit inter granular fracture features and dimples with localized ductile de-
formation and molten metal, which confirm that the projectile material perform typical shear and tensile failure during high-velocity
impact on the rock, accompanied by high temperatures.

Fig. 10. SEM image of the axial fracture surface of fragment 8#.

622
Q. Sun et al. Engineering Failure Analysis 97 (2019) 617–634

Fig. 11. SEM image of the front fracture surface of fragment 8#.

3.2. Projectile structure damage analysis

For thin-walled tubes subjected to axial impact loads, Jones [23] has carried out detailed analyses on the dynamic progressive
buckling and the dynamic plastic buckling with the corresponding structure deformation quite different. He pointed out that the shell
is wrinkled over the entire length when buckled dynamically, unlike the dynamic progressive buckling case with wrinkling confined
to one end (Fig. 12). Apparently, the axial everted fragments collected are split from the projectile cylindrical part, which means the
dynamic plastic buckling other than the dynamic progressive buckling is the main bulking mode occurs during the penetration.
Based on Jones's theory [23] and Chen's analysis [21], three-hinges structures can also be employed to analyze the projectile
plastic buckling. When the projectile nose penetrates into rock partly or totally, the interface between the projectile and the rock
target forms a constraint on the impactor nose which means the degrees of freedom (DOF) of the impactor nose are restricted. Then
the axial inertia effect starts to work as penetration continues. As a result, the dynamic buckling initiates in the projectile transition
part with several hinge structures (red dot in Fig. 13b) approximated at weak position, especially at the section with the thinnest shell
thickness. At the same time, due to the varying shell thickness, the projectile transition part is subjected to the high pressure and the
stress concentration. Afterwards, the projectile transition part expands annularly along radial direction so rapidly that the axial and
lateral inertia effects are forced to respond and then the annular distribution of three-hinges forms (Fig. 13b). Once the three-hinges
formed at the projectile weak positions enter plastic state, the axial carrying capacity of the thin-walled projectile will decrease
sharply and the structural weakening will accelerate. Eventually, fracture along with axial eversion and tearing failure occurs at the
exterior edge of annular expansion. Nevertheless the impactor nose are aslo divided into fragments due to the high stress when the
annular expansion at the transition part fractures. As a result of the cursh of the impactor nose and the transition part, the projectile is
transfromed into a tube. Then the same application of three-hinges structures can analyze the failure of the projectile other parts with
shape lile a tube (Fig. 13c).
Above all, the projectile structure damage with several three-hinges structures employed (Fig. 13c) gives satisfactory explanation

Fig. 12. Two different buckling models [20].

623
Q. Sun et al. Engineering Failure Analysis 97 (2019) 617–634

Fig. 13. Dynamic plastic bulking of the projectile.

for the corresponding experimental results of petal shaped fragments.

4. Numerical simulation

Numerical simulations have become increasingly used for capturing the transient structural response during high-velocity impact
[27]. They provide details of impact process, like stress, strain, temperature and so on, which are valuable to the detailed analysis of
the projectile fracture process.
As a traditional numerical method, Finite Element Method (FEM) is widely applied for the simulation of projectile penetration
[28,29]. However, FEM has one major defect which limits its application on large deformation analysis. It suffers from element
distortion problem that causes negative volume problems and low computational efficiency. Even if the problems are well handled by
element erosion technique and adaptive meshing algorithm, a new problem named of non-conservation of mass comes behind [30].
As a mesh free method, Smoothed Particle Hydrodynamics method (SPH) has been developed to simulate high-velocity impact
problems [30,31], which doesn't cause the calculation problem caused by distortion of elements due to its variable nodal con-
nectivity. More importantly, the mass of whole system keeps unchanged in calculation, which reduces numerical errors caused by the
non-conservation of mass. On account of the computational accuracy and efficiency [30–34], coupled SPH-FEM is also implemented
to reveal the fracture mechanism in the following simulation models.

4.1. Calculation model

According to the geometry schematic of the projectile given in Fig. 1, the model is established in Truegrid-3D and then imported
into the Autodyn software, shown in Fig. 14. The round target is adopted for rock in the simulation with 1200 mm diameter and
800 mm length. In order to reduce the boundary effect in the simulation, the rock target is surrounded by a steel ring with 20 mm
thickness and non-reflect boundary is also applied. Different discrete forms, SPH method (for the whole projectile including shell,
back cover and filling) and FEM (for the rock target including the steel ring), are applied in the simulation model. The particle size is

Fig. 14. Numerical model.

624
Q. Sun et al. Engineering Failure Analysis 97 (2019) 617–634

Table 4
All simulation cases for the projectile with hc = 22mm.
Case vi (m/s) AOA(°) Incidence angle (°) Filling

1 800 0 0 Sulfur
2 900 0 0 Sulfur
3 1000 0 0 Sulfur
4 1100 0 0 Sulfur
5 1200 0 0 Sulfur
3-1 1000 5 0 Sulfur
3-2 1000 10 0 Sulfur
3-3 1000 0 5 Sulfur
3-4 1000 0 10 Sulfur
3-N 1000 0 0 None
3-1-N 1000 5 0 None
3-3-N 1000 0 5 None

2 mm and the total number of particles is about 100,000. The element size in the target center is 4 mm with outward expansion in
proportion and the total number of elements is about 650,000. As the asymmetry property of the random failure for projectile
material during penetration and possible deflection of the projectile, the 1/2 model is adopted and presented in Fig. 14. In addition,
all simulation cases for the projectile with hc = 22mm including different impact velocities, AOA and incidence angles are given in
Table 4.

4.2. Material model

4.2.1. Metal
Since metal material during penetration is subjected to large strains, high rates and high pressures, the Johnson-Cook (JC) model
[35] is selected in the numerical simulation to describe the whole projectile including the shell and the steel ring around rock made
by steel, the back cover made by aluminum. The yield stress equation of JC model is
m
ε̇ T − T0 ⎞ ⎤
σy = (A + Bε n ) ⎛1 + C ln ⎞ ⎡1 − ⎛
⎜ ⎟ ⎜ ⎟

⎝ ε0̇ ⎠ ⎢
⎣ ⎝ Tmelt − T0 ⎠ ⎥
⎦ (1)

where A, B, C, n and m are material constants, σy is the yield stress, ε is the effective strain, ε ̇ is the effective strain rate, ε0̇ is the
reference strain rate, Tmelt is metal melt temperature and Troom is room temperature, T is material transient temperature.
As for the shell material (the bulk material of projetile), 35CrMnSi steel is a very common material with much more research on
its mechanics material parameters for JC model [22,24]. Main material parameters are shown in Table 5.
Previous test results indicate that the projectile is crushed and divided into pieces with different sizes after penetrating the rock
target, so it is necessary to take into account the random failure characteristics of projectile material. The Mott stochastic failure
model [36,37] is an empirical model based on a large number of experimental data, which is widely applicable and can be used
directly in the Autodyn software. The failure probability distribution p of material weakening point in Mott model for arbitrary plastic
strain ε∗ ranges from 0 to 1, expressed as

D ∗
p = 1 − exp ⎜⎛− eγε ⎞⎟
⎝ γ ⎠ (2)

where p is the element failure probability when the proportional plastic strain is ε , D and γ are constant depending on the material
properties. The larger the γ value, the stronger the material uniformity.
In this paper, Mott failure model with the main stress failure is added to the projectile. Specifically, the failure threshold is set as
yield strength (1.5GPa), γ = 10 is selected and the random failure is set from 50% of the failure stress [36,37].

Table 5
Main parameters of JC model for metal material.
Shell Back cover Steel ring

35CrMnSi steel 7039 aluminum 4340 steel

ρ0/(g/cm3) 7.83 2.77 7.83


G/(GPa) 81.8 27.6 81.8
A/(GPa) 1.5 0.337 0.792
B/(GPa) 0.5 0.345 0.51
n 0.26 0.41 0.26
C 0.014 0.01 0.014

625
Q. Sun et al. Engineering Failure Analysis 97 (2019) 617–634

Table 6
Parameters of JH2 model for rock [25].
Parameters Rock

Bulk modulus K1/(GPa) 25.7

Polynomial EOS constant K2/(GPa) −4500


Polynomial EOS constant K3/(GPa 300,000
Shear modulus G/(GPa) 21.9
Hugoniot elastic limit HEL/(GPa) 4.5
Intact strength constant A 0.76
Intact strength exponent N 0.62
Strain rate constant C 0.005
Fractured strength constant B 0.25
Fractured strength exponent M 0.62
Max. fractured strength ratio σf, max∗ 0.25
Hydro tensile limit T/(MPa) −54
Damage constant D1 0.005
Damage constant D2 0.7
Bulking constant β 0.5

4.2.2. Rock
The improved Johnson-Holmquist constitutive (JH-2) model [38], the modified version of original Johnson-Holmquist con-
stitutive [39], was originally formulated for description of brittle response of ceramics, which is implemented as a material model in
AUTODYN hydro-code [3,25,40]. The current material strength of the JH-2 model is:
σ ∗ = σ∗i − D (σ∗i − σ∗f ) (3)

σ∗I = A (P ∗ + T ∗) N (1 + C ln ε ∗̇ ) (3a)

σ∗f = B (P ∗) M [1 + C ln (ε ∗̇ )] (3b)
∗ ∗
where A, B, C, M and N are material constants, D is the damage(ranging from 0 to 1), σ is the normalized equivalent stress, σI is the
normalized intact strength (D = 0), σf∗ is the normalized fractured strength (D = 1), P∗ is the normalized pressure, T∗ is the nor-
malized maximum tensile hydrostatic pressure the material can withstand, ε ∗̇ is the dimensionless strain rate. More detailed de-
scription of JH-2 model can be referred in Ref. [3, 25, 38, 41].
Considering the rock target is a kind of rhyolite having almost the same composition with granite and the study focuses on the
fracture of the projectile other than the target, the JH-2 model parameters for rhyolite in the simulation are referred to those for Barre
Granite with similar physical indices and statics parameters (Table 2) in Ref. [25], as shown in Table 6.

4.2.3. Filling
Sulfur is used as the simulated charge inside the projectile during penetration test. Consulting the Autodyn software material
library [40], only shock equation of state (EOS) is implemented without strength equation, given as:
Us = C1 + S1 Up (4)
where Us is the shock velocity, Up is the particle velocity, S1 is the slope of the Us-Up relationship and C1 is sound speed. Corre-
sponding parameters are shown in Table 7.

4.3. Numerical results

4.3.1. Analysis of projectile fracture mechanism


Fig. 15 presents the process of the projectile penetrating rock at high-velocity (1000 m/s), in which the rock target failure along
with crack propagation (failed elements marked in red color) can be obviously seen. The divergent crack phenomenon on the target
surface in the simulation coincides with that in the field test (Fig. 4).
The failure and fracture process for projectile is presented in Fig. 16. First of all (t = 100 μs in Fig. 16), due to the projectile with
high speed and the rock target with high intensity, the impactor nose surface materials fail severely (failed particles marked in

Table 7
Material parameters of sulfur.
Parameters Sulfur

Density ρ0/(g/cm )
3
2.02
Constant C1/(m/s) 3223
Constant S1 0.95

626
Q. Sun et al. Engineering Failure Analysis 97 (2019) 617–634

Fig. 15. Simulation of penetration into rock (Case 3).

Fig. 16. The projectile shell failure and fracture process (Case 3).

magenta color), which can explain the phenomenon of the impactor nose fractured into small fragments. After removing the failed
particles, the eroded impactor nose is exposed (Fig. 17a). In addition, the corresponding stress contour (Fig. 17b) indicates that the
failed materials on the impactor nose surface is unable to undertake any shear or tensile stress. Based on the material status (Fig. 17a)
and the corresponding stress contour (Fig. 17b), it can also be noted that the projectile transition part (due to its relatively thin shell
thickness) is subjected to the stress concentration at the same time and then fails along with the formation of the plastic yield zone
until it fractures. Furthermore, the sulfur inside with small density and low strength, is seriously extruded by the fragments and
target, which result in that the radial expansion of the projectile transition part and obvious crack at the cylindrical part (t = 300 μs in

627
Q. Sun et al. Engineering Failure Analysis 97 (2019) 617–634

Fig. 17. The material status and corresponding Von-Mises stress contour of the impactor nose at t = 100 μs (Case 3).

Figs. 16 and 18). As the penetration continues, the projectile cylindrical part form everted fragments tearing along the axial direction
(t = 500 μs in Fig. 16). Eventually, at time t = 700 μs (Fig. 15) the tearing damage extends to the projectile tail and then four petals
shaped fragments forms (Fig. 19) obviously. It is worth noting that the split mode of four (or above) petal shaped fragments can also
be observed from the projectile tail fragments collected in the experiment, which means the numerical simulation gives satisfactory
agreement with the corresponding experimental results.
In addition, the shape of the fragment in the calculation is basically consistent with the shape of the last column 9#, 10# and 11#
in Fig. 5 (shown as Fig. 20). While the fourth piece is quite different. This is mainly due to 1/2 reflection model in simulation. The
fully symmetrical fragment shapes will not appear in simulation with full-size models.
Above all, the numerical simulation can effectively reduce the failure process of the projectile penetrating the rock target with
high velocity. By the comparison between the simulation and test results, the Mott stochastic failure model and the material model
parameters adopted in the simulation exhibit high credibility.

4.3.2. The influence of the filling on the projectile fracture mode


In 4.3.1 section, it's mentioned that the fracture of the projectile cylindrical part along axial and radial direction is due to the
expansion of filling (simulated charge). Since the numerical results are not obtained by contrast experiments, the cost of which is
expensive, the contrast simulation is conducted under the same condition except the absence of the filling to verify the proposed
analysis and further reveal the projectile fracture mechanism.
As the contrast simulation, the fracture process of the projectile without filling (Case 3-N) is presented in Fig. 21. It can be found
that, without filling inside, the projectile transforms into approximate cylindrical shell after fractured at the weak section and is
subjected to continuous erosion. None obvious cracks are visible on the projectile surface, which differs from the crack phenomenon
of Case 3 in Fig. 16, especially in the cylindrical part at t = 300 μs. Meanwhile, the corresponding stress contours are performed in
Fig. 22 with the max values of which are almost the same but the stress distribution of which shows obvious differences. The max
values almost the same means the existence of filling or not has few determining influence on the occurrence of the projectile
fracture. While the stress distribution with obvious difference indicates the existence of filling has something significant to do with
the fracture mode. By the way, the max radial distance in case 3 of 282.78 mm is larger than that in case 3-N of 279.58 mm at
t = 300 μs, which indicates the existence of filling accelerate the radial expansion of the projectile transition part.

Fig. 18. Axial crack in projectile shell with different visual angles at t = 300 μs (Case 3).

628
Q. Sun et al. Engineering Failure Analysis 97 (2019) 617–634

Fig. 19. Simulation results of petal shaped fragments at t = 700 μs.

Fig. 20. The comparision between test and numerical results.

Fig. 21. Failure and fracture process of the projectile shell (Case 3-N).

In addition, different final damage modes are shown in Fig. 23, only a short cylinder for the projectile without filling in Case 3-N,
while petal shaped fragments with eversion for the projectiles with filling in Case 3. The same difference can also be found between
the cases with AOA (Case 3-1 and Case 3-1-N) and with incidence angle (Case 3-3 and Case 3-3-N).

629
Q. Sun et al. Engineering Failure Analysis 97 (2019) 617–634

Fig. 22. Different Von-Mises stress contours of the projectile shell at t = 300 μs.

Fig. 23. Final damage modes of the projectiles in different cases.

Apparently, the final damage modes of all contrast simulation cases for the projectiles without filling does not accord with the
corresponding experimental results. The contrast simulations further confirm the existence of filling accelerate the radial expansion
of the projectile transition part and has the determining influence on the projectile damage pattern with the petal shaped fragments.

4.3.3. The influence of the impact conditions on the projectile fracture mode
Projectile final damage modes between different impact velocities are presented in Fig. 24. It can be noted that the crack all
extend to the projectile tail at 800–1200 m/s and the final damage modes are all shown as the split pattern of petal shaped fragments,
consistent with the experimental results. In addition, the higher the velocity the more serious the shell tearing failure with eversion

630
Q. Sun et al.

631
Fig. 24. Final damage modes of the projectiles at different velocites.
Engineering Failure Analysis 97 (2019) 617–634
Q. Sun et al. Engineering Failure Analysis 97 (2019) 617–634

Fig. 25. Final damage modes of the projectiles at different AOA.

and the shorter the projectile rear part. For vi = 1200m/s (Case 5), the projectile are divided into fragments without cylindrical part.
In addition, the number of petal shaped fragments formed indicates no obvious relation with the increase of the velocity, which
indicates the stochastic characteristics of the number of the petal shaped fragments with velocity.
Fig. 25 indicates the projectile final failure pattern at different AOA. With the increase of AOA, the projectile fracture becomes
more asymmetrical and the petal fragment size tends to be more polarized. This failure pattern of polarized petal fragment size is
mainly due to the uneven force when the penetration are carried out with AOA.
Fig. 26 presents the projectile final failure pattern at different incidence angles. The projectile damage modes of the penetration
with different AOA are obvious different from those with different incidence angles. As the incidence angle increases, the numbers of
petal shaped fragments increase while the differences between the fragment sizes decrease and tend to be similar.

5. Conclusion

The penetration experiments into high-strength rock of projectiles with different shell thickness are first conducted at about
1000 m/s. The experimental results indicate that projectiles with different shell thickness are all completely broken and fail to
effectively penetrate the rock target. Besides, the impactor nose divided into small fragments during penetration differs from the
impactor nose almost remaining intact in thin metal target. In addition, the final damage mode of petal shaped fragments with
eversion confirms the projectiles are subjected to the tearing damage and shear failure.
Based on the test results, both factors of material strength failure and projectile structure damage are theoretically discussed to
analyze the projectile failure and fracture mechanism. The SEM technology is applied to confirm the tensile and shear failure oc-
curred during tests in the material strength failure analysis. While the latter factor in the form of the dynamic plastic buckling failure
gives an satisfactory explanation for the fracture pattern of petal shaped fragments with three-hinges structures employed.
Furthermore, the numerical model of penetrating rock target is established by Autodyn-3D. On account of the computational
accuracy and efficiency, coupled SPH-FEM is implemented in the simulation. By taking the failure model of Mott stochastic dis-
tribution into account for the projectiles, numerical simulation can reveal the projectile fracture mechanism satisfactorily and exhibit
high credibility. Additional valuable numerical conclusions concerned with the fracture mechanism of 35CrMnSi steel projectile are
as follows:

1) The serious damage of the impactor nose and the stress concentration forming at the projectile transition part (due to its relatively
thin shell thickness) contributes to the projectile fracture.
2) The final damage mode of the high-velocity projectile penetrating rock is determined by the filling inside. Contrast simulation

632
Q. Sun et al. Engineering Failure Analysis 97 (2019) 617–634

Fig. 26. Final damage modes of the projectiles at different AOA.

results indicate that the split pattern of petal shaped fragments is the final damage mode for the projectiles with filling while a
short cylinder as a result of continues erosion is that for the projectiles without filling.
3) The higher the velocity the more serious the shell tearing failure and the shorter the projectile rear part. Besides, the number of
the petal shaped fragment shows a stochastic characteristics as the velocity increase.
4) With the increase of AOA, the projectile fracture becomes more asymmetrical and the petal fragment size tends to be more
polarized. However, the projectile damage modes of the penetration with different AOA are obviously different from those with
different incidence angles. As the incidence angle increases, the number of petal shaped fragments increases while the differences
between the fragment sizes decrease and tend to be similar.

Above all, the experimental results and the adopted numerical method showing high credibility can be referred to further study of
the projectile fracture mechanism during high-velocity penetration, especially with the emphasis on structure damage analysis.

References

[1] T. Zhan, Z. Wang, J. Ning, Failure behaviors of reinforced concrete beams subjected to high impact loading, Eng. Fail. Anal. 56 (2015) 233–243.
[2] X. Xu, T. Ma, J. Ning, The damage and failure mechanism of the concrete subjected to shaped charge loading, Eng. Fail. Anal. 82 (2017) 741–752.
[3] T. Binar, J. Švarc, P. Vyroubal, T. Kazda, S. Rolc, A. Dvořák, The comparison of numerical simulation of projectile interaction with transparent armour glass for
buildings and vehicles, Eng. Fail. Anal. 92 (2018) 121–139.
[4] F.E. Heuze, An overview of projectile penetration into geological materials, with emphasis on rock, Int. J. Rock Mech. Min. Sci. 27 (1) (1990) 1–14.
[5] C.W. Young, Penetration equations, Office of Scientific & Technical Information, Technical Reports, 33(1−12) 1997, pp. 837–846.
[6] X. Gao, Q.M. Li, Trajectory instability and convergence of the curvilinear motion of a hard projectile in deep penetration, Int. J. Mech. Sci. 121 (2017) 123–142.
[7] Q. Sun, Y. Sun, Y. Liu, R. Li, Y. Zhao, Numerical analysis of the trajectory stability and penetration ability of different lateral-abnormal projectiles for non-normal
penetration into soil based on Modified Integrated Force Law method, Int. J. Impact. Eng. 103 (2017) 159–168.
[8] J. Feng, W. Sun, Z. Liu, C. Cui, X. Wang, An armour-piercing projectile penetration in a double-layered target of ultra-high-performance fiber reinforced concrete
and armour steel: experimental and numerical analyses, Mater. Design 102 (2016) 131–141.
[9] D.Z. Yankelevsky, Resistance of a concrete target to penetration of a rigid projectile – revisited, Int. J. Impact. Eng. 106 (2017) 30–43.
[10] J. Wu, J. Ning, T. Ma, The dynamic response and failure behavior of concrete subjected to new spiral projectile impacts, Eng. Fail. Anal. 79 (2017) 547–564.
[11] R.S. Bernard, Empirical Analysis of Projectile Penetration in Rock, AD-A047989, (1977), pp. 3–20.
[12] A. Darrigade, E. Buzaud, High performance concrete: a numerical and experimental study, 18th International Symposium on Ballistics, 1999, pp. 845–852.
[13] D.J. Frew, M.J. Forrestal, S.J. Hanchak, Penetration experiments with limestone targets and ogive-nose steel projectiles, DE97000740 (1999) 11–12.
[14] D.Z. Zhang, X.R. Zhang, J.D. Lin, R.D. Tang, Penetration experiments for normal impact into granite targets with high-strength steel projectile, Rock. Mech. Eng.
24 (9) (2005) 1612–1618 (in Chinese).
[15] X.Z. Kong, H. Wu, Q. Fang, W. Zhang, Y.K. Xiao, Projectile penetration into mortar targets with a broad range of striking velocities: test and analyses, Int. J.
Impact. Eng. 106 (2017) 18–29.
[16] F. Hu, H. Wu, Q. Fang, J.C. Liu, Impact resistance of concrete targets pre-damaged by explosively formed projectile (EFP) against rigid projectile, Int. J. Impact.

633
Q. Sun et al. Engineering Failure Analysis 97 (2019) 617–634

Eng. 122 (2018) 251–264.


[17] X. Chen, F. Lu, D. Zhang, Penetration trajectory of concrete targets by ogived steel projectiles–experiments and simulations, Int. J. Impact. Eng. 120 (2018)
202–213.
[18] K.G. Rakvåg, T. Børvik, I. Westermann, O.S. Hopperstad, An experimental study on the deformation and fracture modes of steel projectiles during impact, Mater.
Design 51 (5) (2013) 242–256.
[19] K.G. Rakvåg, T. Børvik, O.S. Hopperstad, I. Westermann, Experimental and numerical study on fragmentation of steel projectiles, EPJ. Web Conf. 26 (2012) 1–7.
[20] X. Xiao, W. Zhang, G. Wei, Z. Mu, Effect of projectile hardness on deformation and fracture behavior in the Taylor impact test, Mater. Design 31 (10) (2010)
4913–4920.
[21] X.W. Chen, Y.B. He, G. Chen, M. Qu, Studies on buckling and failure of projectiles under perforation, 23th International Symposium on Ballistics, 2007, pp.
1005–1012.
[22] L. Wang, F.C. Wang, L. Wang, H. Cai, S. Li, Comparative study of penetration performances of steel projectile plates, Acta. Armamentarii. 24 (3) (2003) 419–423
(in Chinese).
[23] N. Jones, Structural Impact, Cambridge University Press, 1997, pp. 390–394.
[24] Q. Li, Dynamic Mechanical Behavior and Fracture Mechanism of 35CrMnSi under Strong Impact Loading, North University of China, 2015, pp. 45–61 (in
Chinese).
[25] M.M.D. Banadaki, B. Mohanty, Numerical simulation of stress wave induced fractures in rock, Int. J. Impact. Eng. 40 (2) (2012) 16–25.
[26] M.A. Meyers, Dynamic Behavior of Materials, John Wiley & Sons, 1994, pp. 350–400.
[27] R. Ranjan, S. Banerji, R.K. Singh, P. Banerji, Local impact effects on concrete target due to missile: an empirical and numerical approach, Ann. Nucl. Energy 68
(2014) 262–275.
[28] T. Børvik, O.S. Hopperstad, T. Berstad, M. Langseth, Perforation of 12 mm thick steel plates by 20 mm diameter projectiles with flat, hemispherical and conical
noses: Part II: numerical simulations, Int. J. Impact. Eng. 27 (1) (2002) 37–64.
[29] S. Dey, T. Børvik, O.S. Hopperstad, J.R. Leinum, M. Langseth, The effect of target strength on the perforation of steel plates using three different projectile nose
shapes, Int. J. Impact. Eng. 30 (8–9) (2004) 1005–1038.
[30] R. Li, Y. Sun, Q. Sun, Y. Zhao, J. Feng, Perforation of steel targets by blunt projectiles using smoothed particle hydrodynamics method, Int. J. Comp. Meth. (2016)
1750044.
[31] G.R. Johnson, R.A. Stryk, S.R. Beissel, SPH for high velocity impact computations, Comput. Method. Appl. M. 139 (1–4) (1996) 347–373.
[32] M.B. Liu, G.R. Liu, Z. Zong, An overview on smoothed particle hydrodynamics, Int. J. Comp. Meth. 5 (1) (2008) 135–188.
[33] Z.S. Liu, S. Swaddiwudhipong, M.J. Islam, Perforation of steel and aluminum targets using a modified Johnson–Cook material model, Nucl. Eng. Des. 250 (2012)
108–115.
[34] S. Swaddiwudhipong, M.J. Islam, Z.S. Liu, High velocity penetration/perforation using coupled smooth particle hydrodynamics-finite element method, Int. J.
Prot. Struct. 1 (2010).
[35] G.R. Johnson, W.H. Cook, Fracture characteristics of three metals subjected to various strains, strain rates, temperatures, and pressures, Eng. Fract. Mech. 21
(1985) 31–48.
[36] J.P. Glanville, G. Fairlie, C. Hayhurst, Numerical Simulation of Fragmentation Using AUTODYN-2D & 3D in Explosive Ordnance Safety Assessment, 6th PARARI
International Explosive Ordnance Symposium, (2003), pp. 29–31.
[37] Y.L. Gui, D.W. Tan, C.W. Sun, 3D simulation of expanding motion and fracture of ductile metal ring, Comput. Mech. 26 (2009) 568–572.
[38] G.R. Johnson, T.J. Holmquist, An Improved Computational Constitutive Model for Brittle Materials, American Institute of Physics, 1994, pp. 981–984.
[39] T.J. Holmquist, G.R. Johnson, W.H. Cook, A computational Constitutive Model for Concrete Subjected to Large Strains, High Strain Rates and High Pressures,
International Symposium on Ballistics, (1993), pp. 591–600.
[40] Autodyn Theory Manual Revision 4.3. Century Dynamics, Concord, CA, (2005), pp. 22–30.
[41] D.S. Cronin, K. Bui, C. Kaufmann, G. McIntosh, T. Berstad, D. Cronin, Implementation and Validation of the Johnson-Holmquist Ceramic Material Model in LS-
DYNA, European Ls, (2004).

634

You might also like