Next Article in Journal
Detecting Multiple Mixed Bacteria Using Dual-Mode Hyperspectral Imaging and Deep Neural Networks
Previous Article in Journal
Numerical and Experimental Characterization of a Coanda-Type Industrial Air Amplifier
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Quasi-Static Finite Cylindrical Cavity Expansion Model for Long-Rod Penetration into Cylindrical Metal Thick Targets with Finite Diameters

Institute of Defense Engineering, Academy of Military Sciences (AMS), PLA, Beijing 100850, China
*
Authors to whom correspondence should be addressed.
Appl. Sci. 2024, 14(4), 1523; https://doi.org/10.3390/app14041523
Submission received: 12 January 2024 / Revised: 31 January 2024 / Accepted: 6 February 2024 / Published: 14 February 2024
(This article belongs to the Section Civil Engineering)

Abstract

:
A quasi-static finite cavity cylindrical expansion model is proposed to investigate the lateral boundary effect during long-rod penetration. Analytical solutions for the cavity pressure in the elastic–plastic stage and plastic stage are obtained, through which a decay function for the lateral boundary effect is constructed. The resistance of the target with finite diameter RT is obtained by multiplying the decay function with the resistance of infinite target RT*. Then, a theoretical model for the penetration into the metal targets with finite diameters by long rods is proposed by substituting RT into the Alekseevskii–Tate model. The proposed model is verified by comparison with existing experimental results. Furthermore, the effect of the target diameter, initial impact velocity and yield criterion on the penetration depth and resistance are investigated. The results showed that the lateral boundary effect should be taken into account if the ratio of the target radius to the rod radius (rt0/RP) is less than 25.

1. Introduction

The problem of long-rod penetration has been extensively researched in the disciplines of protective engineering and armor and weapon design recently [1,2,3]. Unlike rigid projectile penetration, long-rod penetration will cause severe erosion at the contact area between the long-rod and the target, with a semi-fluid deformation mode. In the 1960s, Alekseevskii [4] and Tate [5,6] proposed a theoretical model for the high-speed penetration of long-rods (the Alekseevskii–Tate model). This model assumes that the projectile tail remains rigid during penetration and that erosion occurs only near the contact surface between the projectile and the target. By introducing the projectile strength term YP and the target resistance term RT, the Alekseevskii–Tate model modifies the Bernoulli equation to take into account the strength properties. Subsequent studies have expanded the Alekseevskii–Tate model by considering the effective cross-sectional area [7], transient response stage [8], projectile head shape [9], target response region [10], and target material viscosity [11]. Calculating the appropriate RT is the main focus and challenge in the aforementioned models. Previous studies [6] have demonstrated that utilizing a cavity expansion model to derive RT can yield accurate results.
Bishop [12] first proposed the cavity expansion model, which was primarily used to investigate the issue of normal penetration into semi-infinite targets. Forrestal et al. [13,14] conducted studies on the penetration of rigid projectiles into semi-infinite metal and concrete targets using the cavity expansion model. On this basis, Chen and Li [15] investigated the shape of warheads and proposed a dimensionless head shape function. However, the penetration of finite-diameter targets violates the semi-infinite assumption in the traditional cavity expansion model, thus requiring the introduction of a finite cavity expansion model. Macek et al. [16] first proposed the finite spherical cavity expansion model to study the penetration problem of layered geological materials, taking into account boundary and layering effects.
Recent years have seen the finite cavity expansion model extended to metal [17] and concrete [18] targets, and applied to the study of penetration trajectories [19], perforation [20], and penetration of finite-diameter targets [21,22]. However, the extended model still only applies to rigid projectiles and not to eroding rods directly. Experimental results [23,24] suggest that a long-rod projectile penetrating targets of different diameters at high speed may experience a deviation of approximately 30% in penetration depth. To analyze the influence of the target diameter on high-speed long-rod penetration, Jiang [25] proposed the dynamic finite cylindrical cavity expansion model. It was found that the dynamic resistance is free from the target diameter. He then derived RT based on the average quasi-static cavity pressure instead of the semi-empirical formula in the Alekseevskii–Tate model, thus obtaining the earliest theoretical model of high-speed long-rod penetration into finite-diameter metal targets. However, it is necessary to derive the strength of the target material from the RT obtained from numerical simulations. Furthermore, the use of the dynamic cavity expansion model to calculate quasi-static resistance does not consider the influence of the lateral boundary expansion of the targets. Wang [26] subsequently introduced the unified strength model to analyze different yield criteria. The results showed that different yield criteria can cause a maximum depth change of 15.93%. However, the selection of the yield criterion is dependent on the results of the penetration data.
In summary, while there has been extensive research on the penetration of rigid projectiles, the research on the penetration of long rods into finite-diameter metal targets is insufficient. Considering the displacement of the target, a quasi-static finite cylindrical cavity expansion model for elastic–plastic materials with unified strength theory is proposed and a decay function for the lateral boundary effect is constructed. Then, by combining the decay function and the Alekseevskii–Tate model, an analytical model for the high-speed penetration of finite-diameter metal targets by long rods is established. The model in this paper is verified by experimental and numerical results. Finally, the effects of target diameter, initial impact velocity, and yield criterion on penetration depth and target resistance are examined in detail.

2. Quasi-Static Finite Cylindrical Cavity Expansion Model

In this section, an analytical model for the responses of a finite metal target from the uniform expansion of a cylindrically symmetric cavity is developed. The responses are divided into elastic–plastic and plastic stages based on the size of the cavity radius. Analytical solutions for the cavity pressure are then obtained separately. Based on the solution, the decay function for the lateral boundary effect is obtained using the principle of energy equivalence.

2.1. Basic Model and Equations

It is assumed that, during the penetration process, the target material undergoes only axisymmetric deformation perpendicular to the penetration direction, with no deformation occurring parallel to it. As shown in Figure 1, the radius of the cylindrical target without expansion is rt0, the target at time t is rt, the cavity radius is rc, and the elastic–plastic interface radius is rp. As the cavity radius expands to rc1, the elastic–plastic boundary reaches the target lateral boundary. When the cavity radius expands to rc2, the target fractures. Therefore, the process of cavity expansion can be divided into two stages based on the cavity radius. The first stage is the elastic–plastic stage (rc < rc1), which is composed of the cavity region (0 < r < rc), the plastic region (rc < r < rp), and the elastic region (rp < r < rt). The second stage is the plastic stage (rc1 < rc < rc2), which is composed of the cavity region (0 < r < rc) and the plastic region (rc < r < rt), where r is the radial coordinate in the Euler coordinate (positive outward).
Based on the axisymmetric assumption, the momentum conservation equation in Euler coordinates is
d σ r d r + σ r σ θ r = 0 ,
where σ r and σ θ are the radial and hoop components, respectively, of Cauchy stress taken to be positive under compression.
To obtain a relatively precise analytical solution, it is assumed that the target behaves as an elastic, perfectly plastic rate-independent material according to [15,27]. The target material in the elastic region is assumed to obey Hooke’s law. The stress–displacement relations based on the axisymmetric assumption and Hooke’s law are as follows:
σ r = E ( 1 + v ) ( 1 2 v ) ( 1 v ) d u d r + v u r ,
σ θ = E ( 1 + v ) ( 1 2 v ) v d u d r + ( 1 v ) u r ,
where Yt is the flow stress of the target, E is Young’s modulus of the target material, υ is Poisson’s ratio, and u is the radial displacement (positive outward).
And the unified yield criterion of metal materials under axisymmetric conditions [22] is employed to describe the target material in the plastic region and investigate the resistance characteristics of the target under different yield conditions:
σ r σ θ = 2 ( 1 + b ) 2 + b Y t ,
where Yt is the flow stress of the target and b (0 < b < 1) is the yield criterion coefficient. When b is set to 0, 0.367, and 1, the Tresca yield criterion, Mises yield criterion, and twin shear yield criterion are used, respectively.

2.2. Analytical Solution for Cavity Pressure

2.2.1. Elastic–Plastic Response Stage (rc < rc1)

  • Elastic region (rp < r < rt)
With (2) and (3), (1) transforms to
d 2 u d r 2 + 1 r d u d r u r 2 = 0 .
Since σ r at the lateral boundary of the target is 0 and Equation (4) holds at the elastic–plastic interface, integration gives
u r = ( 1 + b ) ( 1 + v ) Y t ( 2 + b ) E r p 2 r + 1 2 v r p 2 r t 2 r ,
σ r r = ( 1 + b ) Y t ( 2 + b ) r p 2 r 2 r p 2 r t 2 .
  • Plastic region (rc < r < rp)
Substituting Equation (4) into Equation (1) yields
d σ r d r + 2 ( 1 + b ) Y t ( 2 + b ) r = 0 ,
and integration gives
σ r r = ( 1 + b ) Y t ( 2 + b ) 1 r p 2 r t 2 + ln r p 2 r 2 ,
  • Interface Conditions
Assuming that the plastic region is incompressible, it can be concluded that
r p 2 r c 2 = r p u r p 2 .
Substituting Equation (6) into Equation (10) yields
r c 2 + D 2 r p 2 + 2 ( 1 2 v ) r p 4 r t 2 + ( 1 2 v ) 2 r p 6 r t 4 = 2 D ( 1 2 v ) r p 4 r t 2 + 2 D r p 2 ,
D = ( 1 + b ) ( 1 + v ) Y t ( 2 + b ) E .
With u r t = r t r t 0 , (6) transforms to
2 ( 1 + b ) ( 1 v 2 ) Y t ( 2 + b ) E r p 2 = r t 2 r t 0 r t .
If rc = rc1, then rp = rt. From (10), and ignoring higher order terms in u(rp) according to [28],
r c 1 2 = 2 u r t 1 r t 1 ,
where rt1 is the target radius if rc = rc1. From (6), (13), and (14), rc1 can be obtained
r c 1 = 2 ( 1 + b ) ( 2 + b ) ( 1 v 2 ) E Y t ( 2 + b ) E 2 ( 1 + b ) ( 1 v 2 ) Y t r t 0 .

2.2.2. Plastic Response Stage (rc1 < rc < rc2)

With the lateral boundary conditions, integration of (8) gives
σ r r = ( 1 + b ) Y t ( 2 + b ) ln r t 2 r 2 .
It can be concluded from (10) that
r t 2 = r t 0 2 + r c 2 .
Substituting Equation (17) into Equation (16) yields
σ r r = ( 1 + b ) Y t ( 2 + b ) ln r t 0 2 + r c 2 r 2 .
rc2 can be taken as [26]
r c 2 = r t 2 + b ε f 1 + b ,
where εf is the uniaxial ultimate strain. Substituting Equation (17) into Equation (19) gives rc2 considering the deformation of the target:
r c 2 = 2 + b ε f 1 + b 1 2 + b ε f 1 + b r t 0 .

2.2.3. Semi-Infinite Target

For a semi-infinite target ( r t ), σ r can be obtained from Equations (9) and (11):
σ r r = ( 1 + b ) Y t ( 2 + b ) 1 + ln r c 2 r 2 ( 2 + b ) 2 E 2 2 ( 1 + b ) ( 2 + b ) ( 1 + v ) E Y t ( 1 + b ) ( 1 + v ) Y t 2 .
In summary, the cavity stress at each stage for the finite target and the semi-infinite target is
σ r c = ( 1 + b ) Y t ( 2 + b ) 1 r p 2 r t 2 + ln r p 2 r c 2 , r c < r c 1 ( 1 + b ) Y t ( 2 + b ) ln 1 + r t 0 2 r c 2 , r c 1 r c r c 2 0 , r c r c 2 ( 1 + b ) Y t ( 2 + b ) 1 + ln r c 2 r 2 ( 2 + b ) 2 E 2 2 ( 1 + b ) ( 2 + b ) ( 1 + v ) E Y t ( 1 + b ) ( 1 + v ) Y t 2 , r t .
where σ r c is the radial stress at the cavity.

2.3. Average Cavity Pressure and the Decay Function for the Lateral Boundary Effect

The energy dissipated during the expansion of the cavity radius from 0 to the final value rcf is equal to the work performed by the cavity pressure [25]. Therefore, the energy dissipated per unit thickness is
W e = 0 r c f 2 π r c σ r c d r c = 2 π 0 r c f σ r c r c d r c ,
assuming that the work performed by the average pressure of the cavity σ ¯ r c on the displacement of the expansion is equal to the energy dissipated of the expansion, σ ¯ r c can be obtained from the conservation of energy as follows:
σ ¯ r c = W e π r c f 2 = 0 r c f σ r c r c d r c 1 2 r c f 2 .
With (22) and (24), the decay function for the lateral boundary effect can be given as follows:
f r c f , r t 0 = σ ¯ r c σ r c r t .

3. Theoretical Penetration Model and Model Verification

In this section, a modified Alekseevskii–Tate model based on the decay function for the lateral boundary effect is developed. Analytical solutions for the penetration depth are obtained. The penetration depth results from existing experiments and numerical simulations are subsequently used to validate the proposed model.

3.1. Modified Alekseevskii–Tate Model and Calculation of Penetration Depth

According to Tate [5,6], it is assumed that the interface of the high-velocity long-rod and the target is semi-fluid and the initial shape of the long-rod’s nose has negligible effects on the penetration depth. As shown in Figure 2, RP is the radius of the rod; v0 and L0 are the initial impact velocity and the initial length of the projectile, respectively; and u, v, P, and L are the penetration velocity, tail velocity, penetration depth, and remaining length of the rod, respectively. The following basic equations can be drawn according to the Alekseevskii–Tate model [5]:
1 2 ρ P v u 2 + Y P = 1 2 ρ T u 2 + R T ,
d l d t = v u ,
d P d t = u ,
d v d t = Y P ρ P L .
The initial conditions are as follows:
P ( 0 ) = 0 , L ( 0 ) = L 0 , v ( 0 ) = v 0 ,
where ρ P and ρ T are the density of the rod and target, respectively; YP is the dynamic flow stress of the rod; and RT is the target resistance.
YP can be taken as [29]
Y P = ( 1 + λ ) σ yp .
By using (25), RT can be modified as
R T = f r c f , r t 0 R T * ,
R T * = Y t 2 3 + ln 2 E 4 e λ Y t ,
where σ yp is the yield stress of the rod; λ is usually taken as 0.7; Rt* is the target resistance of a semi-infinite target [29]; and rcf can be taken as the channel radius [8]:
r c f = R P 1 + 0.287 v 0 + 0.148 v 0 2 ,
where v0 is measured in km/s.
Using the above equations, the depth of penetration can be calculated in two situations determined by the relationship between YP and Rt.
  • YP < Rt
In this situation, the long-rod only undergoes eroding penetration. When v = vc and u = 0 are reached, the penetration ends, and the remaining long-rod continues to erode until the residual velocity reaches 0. The penetration depth can be obtained as follows:
P = ρ P L 0 Y P v c v 0 1 1 λ 2 v λ v 2 + A υ ( v ) υ ( v 0 ) k 1 exp k 2 [ ϑ ( v ) ϑ ( v 0 ) ] d v ,
v c = 2 ( R T Y P ) ρ P , λ = ρ T ρ P , A = 2 ( R T Y P ) ρ T 1 λ 2 ,
υ x = x + x 2 + A , ϑ x = x x 2 + A λ x 2 ,
k 1 = R T Y P λ Y P , k 2 = λ ρ P 2 1 λ 2 Y P .
  • YP > Rt
In this situation, the penetration depth can be divided into two parts: eroding penetration and rigid body penetration. First, the high-speed long-rod undergoes eroding penetration until the penetration speed is equal to the tail velocity (u = v = vr). Then, the remaining long-rod (with remaining length Lr) penetrates as a rigid body at the remaining velocity vr. The depth of penetration can therefore be obtained as
P 1 = ρ P L 0 Y P v r v 0 1 1 μ 2 v μ v 2 + A υ ( v ) υ ( v 0 ) k 1 exp k 2 [ ϑ ( v ) ϑ ( v 0 ) ] d v ,
v r = 2 ( Y P R T ) ρ T ,
P 2 = ρ P L r R P 2 v r 2 2 r c f 2 R T ,
r c f = R P 1 + 0.287 v 0 + 0.148 v 0 2 ,
L r = L 0 υ v r υ v 0 k 1 exp k 2 ϑ v r ϑ v 0 ,
P = P 1 + P 2 .

3.2. Validation of the Theoretical Model

In Section 3.2, b = 0.367, i.e., the Mises yield criterion, is used in the proposed model. Experiments on long rods with different initial impact velocities penetrating metal targets with different diameters were carried out in [23,24]. The mechanical parameters of the rod and target materials are summarized in Table 1. As the flow stress is not always given, it can be estimated from the Brinell hardness [30]:
σ y = 3.92 × BHN ,
where σ y is the flow stress, in MPa; BHN is the Brinell hardness.
It is clear from the figure in [23] that the target undergoes significant radial expansion after penetration, which is consistent with the model in this article. In addition, numerical simulation calculations show that the elastic–plastic interface just reaches the lateral boundary of the target when rt0/RP is approximately 15 [23], while the theoretical model in this paper and other references predicts that the elastic–plastic interface reaches the lateral boundary of the target when rt0/RP = 16.63, 19.92 [23], 19.3 [25], and 17.4 [26]. The proposed model is closest to the numerical simulation results.
Figure 3 shows the comparison of experimental and numerical results [23] with the calculation results of penetration depths for different rt0/RP. Table 2 and Table 3 summarize the initial penetration conditions, penetration depth results, and predicted penetration depths of different theoretical models in [23,24], respectively. From Figure 3 and Table 2, the following conclusions can be drawn:
  • The proposed model is in good agreement with the experimental and numerical results, while all the other theoretical models overestimate the penetration depth;
  • Although Wang’s [26] model has better prediction results than Jiang’s [25] and Littlefield’s [23] models, there is a sudden change in the penetration depth when rt0/RP is around 18.5, which will result in multiple results within a certain range of rt0/RP.
Let the ratio of the resistance of a finite-diameter target to the resistance of a semi-infinite target RT/RT* be the normalized target resistance. Figure 4 shows the comparison of the calculated normalized target resistance with the numerical results [24]. From Figure 4, the following can be concluded:
  • The proposed model is in good agreement with the numerical results. The calculated results are smaller than the numerical results when rt0/RP > 6.5. The reason is that the model does not consider the propagation time of elastic waves during the elastic–plastic stage;
  • Due to the assumption that the target material is incompressible during the plastic stage, the target resistance is overestimated when rt0/RP < 6.5 compared to the numerical results.
The comparison of the calculation results of penetration depths with experimental results [24] for different initial impact velocities is shown in Figure 5. From Figure 5, the following can be concluded:
  • The proposed model is in good agreement with the experimental results. The calculation results of the proposed model continue to increase and tend towards the fluid dynamic limit as the initial impact velocity increases;
  • The calculation results of the other three models are all significantly greater than the experimental results because the RT predicted by the three models is less than YP. When the penetration velocity is less than vr ≈ 2500 m/s, there is only a rigid body penetration phase. However, when the initial impact velocity is greater than vr, the penetration process exhibits an eroding penetration stage. The calculation results of the three models rapidly decrease and tend towards the fluid dynamic limit.
Table 4 shows the mean percentage error, root mean square error, and mean absolute error of the theoretical models. The results show that the proposed model has higher accuracy and better applicability.

4. Results and Discussions

In this section, the effects of the target diameter, initial impact velocity, and yield criterion are investigated based on the material parameters in [23].

4.1. Effect of Target Diameter

Figure 6 shows the comparison of normalized target resistance and penetration depth with rt0/RP at initial impact velocities of 1000 m/s, 1500 m/s, and 2000 m/s. Figure 7 shows the comparison of penetration velocity and tail velocity with penetration depth at different rt0/RP values of 5, 10, and 20. From Figure 6 and Figure 7, the following can be seen:
  • When rt0/RP < 25, Rt/Rt* decreases rapidly as rt0/RP decreases, while Rt/Rt* decreases slowly as rt0/RP decreases when rt0/RP > 25. This suggests that Rt/Rt* rapidly decreases as the elastic–plastic interface approaches the target boundary (rt0/RP = 16.63).
  • When rt0/RP < 25, the penetration depth increases exponentially with decreasing rt0/RP. However, when rt0/RP > 25, the penetration depth slowly increases with decreasing rt0/RP. This is because of the decrease in Rt/Rt*. When rt0/RP = 25, the deviation of the penetration depth from the semi-infinite target is less than 10%, which means that the target can be treated as a semi-infinite target when rt0/RP > 25.
  • As the circumferential constraint effect decreases, the initial penetration velocity and penetration depth gradually increase.
Figure 8 compares the decay functions of the target resistance in this article with those in [17]. As shown in Figure 8, the lateral boundary effect in the elastic–plastic stage is relatively small. When the elastic–plastic interface approaches the lateral boundary, the target resistance decreases by about 13%. Different from the decay functions in [17], the following can be seen:
  • Because the proposed decay function is based on the average cavity pressure rather than the instantaneous cavity pressure, the proposed decay function decreases more slowly than that in [17];
  • The diameter of the target is greater as the elastic–plastic interface intersects the target boundary in the proposed decay function because of the use of the finite cylindrical cavity expansion model;
  • The proposed decay function will not decrease to 0 before the initial target radius decreases to 0. The reason for this is that the target deformation is taken into account in this article;
  • Due to the fact that the target resistance of long-rod penetration is independent of velocity, the decay function in this article is independent of velocity, whereas the decay function in [17] decreases significantly with increasing velocity.

4.2. Effect of Initial Impact Velocity

Figure 9 shows the comparison of normalized target resistance and penetration depth with initial impact velocity at rt0/RP values of 5, 10, and 20. Figure 10 shows the comparison of penetration velocity and tail velocity with penetration depth at initial impact velocities of 1000 m/s, 1500 m/s, and 2000 m/s. From Figure 9 and Figure 10, the following can be seen:
  • The penetration depth increases with increasing initial impact velocity and gradually approaches the fluid dynamic limit. In the case of rt0/RP = 5, when the initial impact velocity is greater than 2500 m/s, the penetration depth will be slightly greater than the fluid dynamic limit. This is because when v0 > 2500 m/s at rt0/RP = 5, RT is less than YP, and rigid penetration occurs during the penetration process;
  • Rt/Rt* decreases approximately linearly as the initial impact velocity increases. This is because rcf increases as the initial impact velocity increases, leading to an increase in the lateral boundary effect;
  • Both the initial penetration velocity and penetration depth increase significantly as the initial impact velocity increases.

4.3. Effect of Yield Criteria

With different b values, different yield criteria can be applied in the proposed model. Figure 11 shows the variation in penetration depth and normalized target resistance with rt0/RP when the Tresca yield criterion (b = 0), twin shear yield criterion (b = 1), and Mises yield criterion (b = 0.367) are used, respectively. Table 5 shows the ranges of penetration depth and normalized target resistance at different rt0/RP. Figure 11 and Table 5 show the following:
  • As b increases, the penetration depth decreases and Rt/Rt* increases slightly;
  • The calculated penetration depth and target resistance are less affected by the yield criterion. For different yield criteria, the maximum deviations in the penetration depth and penetration resistance are 2.36% and 4.62%, respectively. The reason is that the Rt* used in this model is independent of the yield criterion.
Figure 11. Comparison of penetration depth and normalized target resistance of different yield criteria: (a) penetration depth; (b) normalized target resistance.
Figure 11. Comparison of penetration depth and normalized target resistance of different yield criteria: (a) penetration depth; (b) normalized target resistance.
Applsci 14 01523 g011
Table 5. Summary of penetration depth and normalized target resistance ranges.
Table 5. Summary of penetration depth and normalized target resistance ranges.
rt0/RPRange of Penetration Depths (mm)Range of Rt/Rt*
3114.74~117.440.368~0.386
595.20~97.740.519~0.543
785.37~87.770.622~0.649
979.23~81.480.697~0.726
1175.04~77.110.754~0.783
1372.05~73.910.799~0.827
1569.89~71.520.835~0.860
1768.24~69.710.863~0.886
1967.09~68.270.886~0.905
2166.21~67.230.903~0.920
3064.05~64.620.947~0.957
8062.04~62.120.991~0.993

5. Summary and Conclusions

In summary, a theoretical model of high-speed penetration of finite-diameter metal targets by a long-rod is established using a quasi-static finite cylindrical cavity expansion model. The calculated results are in good agreement with the experimental results. Studies have also been conducted on the factors influencing the long-rod penetration into metal targets with finite diameters. The conclusions are as follows:
  • The proposed model requires only the material and size parameters of the target and the long-rod as inputs, and predicts the penetration depth well across a range of rt0/RP, varying target and long-rod materials, and diverse v0;
  • The penetration depth increases by less than 10% between semi-infinite targets and rt0/RP = 25. However, an increase of more than 50% in the penetration depth occurs between rt0/RP = 25 and rt0/RP = 5. Therefore, the lateral boundary effect should be taken into account when rt0/RP < 25;
  • The penetration depth increases dramatically with increasing v0 before the fluid dynamic limit. On the other hand, Rt decreases significantly with v0 due to the increase in the channel radius caused by v0. Therefore, the lateral boundary effect is also influenced by v0;
  • The penetration depth decreases very little as b increases, which indicates that the yield criterion has little effect on the lateral boundary effect and Rt.

Author Contributions

Conceptualization, Y.L. and X.T.; methodology, Y.L.; software, Y.L.; validation, Y.L. and J.S.; formal analysis, Y.L.; investigation, Y.L. and B.Z.; resources, X.T.; data curation, Y.L., X.X. and J.S.; writing—original draft preparation, Y.L.; writing—review and editing, X.T., Y.C. and B.Z.; visualization, Y.L., X.X. and J.W.; supervision, X.T.; funding acquisition, X.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Acknowledgments

Yunke Lu would like to extend his sincere gratitude to Cong Sun for her invaluable support.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Orphal, D.L. Explosions and Impacts. Int. J. Impact Eng. 2006, 33, 496–545. [Google Scholar] [CrossRef]
  2. Rosenberg, Z.; Dekel, E. Terminal Ballistics; Springer: Singapore, 2016; pp. 1–76. ISBN 978-981-10-0393-6. [Google Scholar]
  3. Anderson, C.E., Jr. Analytical Models for Penetration Mechanics: A Review. Int. J. Impact Eng. 2017, 108, 3–26. [Google Scholar] [CrossRef]
  4. Alekseevskii, V.P. Penetration of a Rod into a Target at High Velocity. Combust. Explos. Shock Waves 1966, 2, 63–66. [Google Scholar] [CrossRef]
  5. Tate, A. A Theory for the Deceleration of Long Rods after Impact. J. Mech. Phys. Solids 1967, 15, 387–399. [Google Scholar] [CrossRef]
  6. Tate, A. Further Results in the Theory of Long Rod Penetration. J. Mech. Phys. Solids 1969, 17, 141–150. [Google Scholar] [CrossRef]
  7. Rosenberg, Z.; Marmor, E.; Mayseless, M. On the Hydrodynamic Theory of Long-Rod Penetration. Int. J. Impact Eng. 1990, 10, 483–486. [Google Scholar] [CrossRef]
  8. Walker, J.D.; Anderson, C.E. A Time-Dependent Model for Long-Rod Penetration. Int. J. Impact Eng. 1995, 16, 19–48. [Google Scholar] [CrossRef]
  9. Zhang, L.S.; Huang, F.L. Model for Long-Rod Penetration into Semi-Infinite Targets. J. Beijing Inst. Technol. 2004, 13, 285–289. [Google Scholar] [CrossRef]
  10. Lan, B.; Wen, H. Alekseevskii-Tate Revisited: An Extension to the Modified Hydrodynamic Theory of Long Rod Penetration. Sci. China Technol. Sci. 2010, 53, 1364–1373. [Google Scholar] [CrossRef]
  11. Kong, X.Z.; Wu, H.; Fang, Q.; Peng, Y. Rigid and Eroding Projectile Penetration into Concrete Targets Based on an Extended Dynamic Cavity Expansion Model. Int. J. Impact Eng. 2017, 100, 13–22. [Google Scholar] [CrossRef]
  12. Bishop, R.F.; Hill, R.; Mott, N.F. The Theory of Indentation and Hardness Tests. Proc. Phys. Soc. 1945, 57, 147. [Google Scholar] [CrossRef]
  13. Warren, T.L.; Forrestal, M.J. Effects of Strain Hardening and Strain-Rate Sensitivity on the Penetration of Aluminum Targets with Spherical-Nosed Rods. Int. J. Solids Struct. 1998, 35, 3737–3753. [Google Scholar] [CrossRef]
  14. Forrestal, M.J.; Tzou, D.Y. A Spherical Cavity-Expansion Penetration Model for Concrete Targets. Int. J. Solids Struct. 1997, 34, 4127–4146. [Google Scholar] [CrossRef]
  15. Chen, X.W.; Li, Q.M. Deep Penetration of a Non-Deformable Projectile with Different Geometrical Characteristics. Int. J. Impact Eng. 2002, 27, 619–637. [Google Scholar] [CrossRef]
  16. Macek, R.W.; Duffey, T.A. Finite Cavity Expansion Method for Near-Surface Effects and Layering during Earth Penetration. Int. J. Impact Eng. 2000, 24, 239–258. [Google Scholar] [CrossRef]
  17. Warren, T.L.; Poormon, K.L. Penetration of 6061-T6511 Aluminum Targets by Ogive-Nosed VAR 4340 Steel Projectiles at Oblique Angles: Experiments and Simulations. Int. J. Impact Eng. 2001, 25, 993–1022. [Google Scholar] [CrossRef]
  18. Warren, T.L.; Hanchak, S.J.; Poormon, K.L. Penetration of Limestone Targets by Ogive-Nosed VAR 4340 Steel Projectiles at Oblique Angles: Experiments and Simulations. Int. J. Impact Eng. 2004, 30, 1307–1331. [Google Scholar] [CrossRef]
  19. Chen, X.; Lu, F.; Zhang, D. Penetration Trajectory of Concrete Targets by Ogived Steel Projectiles–Experiments and Simulations. Int. J. Impact Eng. 2018, 120, 202–213. [Google Scholar] [CrossRef]
  20. Wei, H.; Liu, J.; Zhang, X.; Liu, C.; Li, P.; Qiao, L. Study on Perforation of Elliptical Cross-Section Projectile into Finite-Thick Metal Targets. Acta Mech. Sin. 2023, 39, 422429. [Google Scholar] [CrossRef]
  21. Song, D.; Liu, F.; Jiang, Z. Prediction of Projectile an Analytical Model for Penetration into Cylindrical Metallic Thick Target by Rigid Sharp-Nosed Projectiles. Eng. Mech. 2013, 30, 31–36. (In Chinese) [Google Scholar] [CrossRef]
  22. Wang, J.; Zhao, J.; Zhou, Y.; Sun, S. Boundary Effect and Dynamic Response Study for the Penetration of Rigid Projectiles into Thick, Finite-Radius, Metallic Targets. Shock Vib. 2020, 2020, e8832925. [Google Scholar] [CrossRef]
  23. Littlefield, D.L.; Anderson, C.E.; Partom, Y.; Bless, S.J. The Penetration of Steel Targets Finite in Radial Extent. Int. J. Impact Eng. 1997, 19, 49–62. [Google Scholar] [CrossRef]
  24. Forrestal, M.J.; Piekutowski, A.J. Penetration Experiments with 6061-T6511 Aluminum Targets and Spherical-Nose Steel Projectiles at Striking Velocities between 0.5 and 3.0km/s. Int. J. Impact Eng. 2000, 24, 57–67. [Google Scholar] [CrossRef]
  25. Jiang, Z.; Song, D.; Zeng, S. Prediction of Projectile—A Finite Cylindrical Cavity Expansion Theory and Its Application. J. Vib. Shock 2011, 30, 139–143+204. (In Chinese) [Google Scholar] [CrossRef]
  26. Wang, J.; Zhao, J.; Zhang, J.; Zhou, Y. Study on Model of Penetration into Thick Metallic Targets with Finite Planar Sizes by Long Rods. Shock Vib. 2021, 2021, e6666770. [Google Scholar] [CrossRef]
  27. Forrestal, M.J.; Okajima, K.; Luk, V.K. Penetration of 6061-T651 Aluminum Targets with Rigid Long Rods. J. Appl. Mech. 1988, 55, 755–760. [Google Scholar] [CrossRef]
  28. Satapathy, S.; Bless, S. Calculation of Penetration Resistance of Brittle Materials Using Spherical Cavity Expansion Analysis. Mech. Mater. 1996, 23, 323–330. [Google Scholar] [CrossRef]
  29. Tate, A. Long Rod Penetration Models—Part II. Extensions to the Hydrodynamic Theory of Penetration. Int. J. Mech. Sci. 1986, 28, 599–612. [Google Scholar] [CrossRef]
  30. Recht, R.F. Taylor Ballistic Impact Modelling Applied to Deformation and Mass Loss Determinations. Int. J. Eng. Sci. 1978, 16, 809–827. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of quasi-static finite cylindrical cavity expansion model: (a) elastic–plastic stage; (b) plastic stage.
Figure 1. Schematic diagram of quasi-static finite cylindrical cavity expansion model: (a) elastic–plastic stage; (b) plastic stage.
Applsci 14 01523 g001
Figure 2. Schematic diagram of long-rod and cylindrical metal thick target with finite diameter: (a) before impact; (b) shortly after impact.
Figure 2. Schematic diagram of long-rod and cylindrical metal thick target with finite diameter: (a) before impact; (b) shortly after impact.
Applsci 14 01523 g002
Figure 3. Comparison of calculation results of penetration depths using the model in [23,25,26] and in this article with experimental and numerical results in [23].
Figure 3. Comparison of calculation results of penetration depths using the model in [23,25,26] and in this article with experimental and numerical results in [23].
Applsci 14 01523 g003
Figure 4. Comparison of results of the normalized target resistance with numerical results in [23].
Figure 4. Comparison of results of the normalized target resistance with numerical results in [23].
Applsci 14 01523 g004
Figure 5. Comparison of calculation results of penetration depths using the model in [23,25,26] and in this article with existing experimental results in [24].
Figure 5. Comparison of calculation results of penetration depths using the model in [23,25,26] and in this article with existing experimental results in [24].
Applsci 14 01523 g005
Figure 6. Comparison of normalized target resistance and penetration depth with rt0/RP.
Figure 6. Comparison of normalized target resistance and penetration depth with rt0/RP.
Applsci 14 01523 g006
Figure 7. Comparison of penetration velocity and tail velocity with penetration depth of different rt0/RP.
Figure 7. Comparison of penetration velocity and tail velocity with penetration depth of different rt0/RP.
Applsci 14 01523 g007
Figure 8. Decay functions in this article and [17] vs. rt0/rc.
Figure 8. Decay functions in this article and [17] vs. rt0/rc.
Applsci 14 01523 g008
Figure 9. Comparison of normalized target resistance and penetration depth with initial impact velocity.
Figure 9. Comparison of normalized target resistance and penetration depth with initial impact velocity.
Applsci 14 01523 g009
Figure 10. Comparison of penetration velocity and tail velocity with penetration depth of different initial impact velocities.
Figure 10. Comparison of penetration velocity and tail velocity with penetration depth of different initial impact velocities.
Applsci 14 01523 g010
Table 1. Mechanical parameters of long-rod and target material in references.
Table 1. Mechanical parameters of long-rod and target material in references.
Referenceσyp (MPa)Yt (MPa)ρP (kg/m3)ρT (kg/m3)L0 (mm)RP (mm)
Littlefield et al. [23]13001055 117,730785077.93.895
Forrestal et al. [24]1140 [13]400 [13]7830271074.6553.555
1 Calculated using Equation (45).
Table 2. Summary of initial penetration conditions, penetration depths, and calculation results of different theoretical model [23].
Table 2. Summary of initial penetration conditions, penetration depths, and calculation results of different theoretical model [23].
Resourcev0 (m/s)rt0/RPPe (mm) *Pmax (mm)Pmax2 [25] (mm)Pmax3 [26] (mm)Pmax4 [23] (mm)
Experimental results [23]15004.9087.2597.14105.59102.96105.65
15004.9088.4997.14105.59102.96105.65
15006.5386.3188.4799.4996.4999.90
15006.5286.0088.5199.5296.5299.93
150013.1070.9772.8387.7283.1488.47
150013.1274.8672.8187.7083.1288.45
150019.8369.2567.2383.1180.6083.97
150019.7969.7267.2583.1380.6283.98
150019.8866.7667.2083.0980.5883.95
Numerical results [23]15003.26105.94112.38115.69113.51114.64
15004.8989.6697.20105.64103.01105.69
15006.5284.3788.5199.5296.5299.93
150013.0469.8872.9187.7883.2288.53
150019.5667.6267.3683.2280.7184.08
150029.3467.1564.4380.9478.5281.57
150077.6666.8462.1079.4277.0679.55
* Pe are the experimental and numerical results of penetration depth in [23]; Pmax, Pmax2, Pmax3, and Pmax4 are penetration depths calculated by the proposed model and the models in [23,25,26], respectively.
Table 3. Summary of initial penetration conditions, penetration depths, and calculation results of different theoretical models [24].
Table 3. Summary of initial penetration conditions, penetration depths, and calculation results of different theoretical models [24].
v0 (m/s)rt0/RPPe (mm) *Pmax (mm)Pmax2 [25] (mm)Pmax3 [26] (mm)Pmax4 [23] (mm)
103735.1664.657.44116.52115.53117.07
104235.1641.657.90117.17116.17117.72
117435.1667.569.61133.54132.40134.11
117435.1666.569.61133.54132.40134.11
119335.1650.771.22135.77134.61136.33
121635.1650.773.14138.42137.24138.98
128435.1678.878.61145.94144.70146.49
133735.1661.882.64151.48150.19152.01
141135.16106.187.91158.73157.38159.23
151535.167694.58167.96166.53168.40
180235.1694.3108.50187.81186.20187.97
181335.16120108.92188.41186.80188.56
205235.16113.9116.01199.00197.29198.80
220435.16124.6119.00203.43201.68202.96
225535.16137.4119.80204.57202.81204.00
247635.16137.9122.34207.70205.92206.66
* Pe are the experimental results of penetration depth in [24]; Pmax, Pmax2, Pmax3, and Pmax4 are penetration depths calculated by the proposed model and the models in [23,25,26], respectively.
Table 4. Summary of statistical errors.
Table 4. Summary of statistical errors.
MethodStatistical ErrorExperimental Results [23]Numerical Results [23]Experimental Results [24]
The present modelMPE *2.29%1.75%9.47%
RMSE3.564.1211.78
MAE0.060.060.22
Jiang [25]MPE19.53%19.01%94.73%
RMSE15.0314.3971.60
MAE0.200.201.07
Wang [26]MPE15.48%15.30%93.12%
RMSE11.9311.5970.28
MAE0.160.161.06
Littlefield [23]MPE20.30%19.44%95.13%
RMSE15.5914.6571.78
MAE0.210.201.08
* MPE, RMSE, and MAE are mean percentage error, root mean square error, and mean absolute error, respectively.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Lu, Y.; Tao, X.; Chen, Y.; Zhang, B.; Song, J.; Xu, X.; Wang, J. Quasi-Static Finite Cylindrical Cavity Expansion Model for Long-Rod Penetration into Cylindrical Metal Thick Targets with Finite Diameters. Appl. Sci. 2024, 14, 1523. https://doi.org/10.3390/app14041523

AMA Style

Lu Y, Tao X, Chen Y, Zhang B, Song J, Xu X, Wang J. Quasi-Static Finite Cylindrical Cavity Expansion Model for Long-Rod Penetration into Cylindrical Metal Thick Targets with Finite Diameters. Applied Sciences. 2024; 14(4):1523. https://doi.org/10.3390/app14041523

Chicago/Turabian Style

Lu, Yunke, Xigui Tao, Yicun Chen, Bei Zhang, Jiageng Song, Xiangyu Xu, and Jianshuai Wang. 2024. "Quasi-Static Finite Cylindrical Cavity Expansion Model for Long-Rod Penetration into Cylindrical Metal Thick Targets with Finite Diameters" Applied Sciences 14, no. 4: 1523. https://doi.org/10.3390/app14041523

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop