Log in

Oblique Penetration of Spherical Projectile into Low-Carbon Steel Target: Experiment, Theory, 3D Penetration Model

  • Published:
Mechanics of Solids Aims and scope Submit manuscript

Abstract

To solve the ballistic limit and trajectory deflection angle of a spherical projectile after oblique penetration of a finite-thickness mild steel plate, a ballistic penetration model and its construction method are proposed, which combine the theory of cavity expansion penetration, the practice of projectile and target separation modeling, the technique of shooting line projectile-target intersection. The critical algorithms and calculation steps required to construct the penetration model in three-dimensional space are systematically described. Subsequently, the experiments of 93W spherical projectiles with 6mm diameter penetrating 4, 6, and 8 mm Q345 steel targets at angles of 0°, 20°, and 40° are carried, and obtain the penetration ballistic limit and trajectory. Finally, according to the experiment results, the calculation accuracy of the model is checked, and the results show that the ballistic limit calculation maximum errors for 4, 6, and 8 mm Q345 steel targets are 6.69, 18.24, and 16%. The main work of this paper shows that the penetration model established in this paper can accurately calculate the problem of spherical projectile penetrating finite thickness targets and can provide a new solution method for the problem of projectile penetration.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Subscribe and save

Springer+ Basic
EUR 32.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or Ebook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

Price includes VAT (Germany)

Instant access to the full article PDF.

Fig. 1.
Fig. 2.
Fig. 3.
Fig. 4.
Fig. 5.
Fig. 6.
Fig. 7.
Fig. 8.
Fig. 9.
Fig. 10.
Fig. 11.
Fig. 12.
Fig. 13.
Fig. 14.
Fig. 15.
Fig. 16.

REFERENCES

  1. M. E. Backman and W. Goldsmith, “The mechanics of penetration of projectiles into targets,” Int. J. Eng Sci. 16 (1), 1–99 (1978). https://doi.org/10.1016/0020-7225(78)90002-2

    Article  Google Scholar 

  2. C. E. Anderson and J. J. P. Riegel, “A penetration model for metallic targets based on experimental data,” Int. J. Impact Eng. 80, 24–35 (2015). https://doi.org/10.1016/j.ijimpeng.2014.12.009

    Article  Google Scholar 

  3. W. Schonberg and S. Ryan, “Predicting metallic armour performance when impacted by fragment-simulating projectiles – model review and assessment,” Int. J. Impact Eng. 158, 104025 (2021). https://doi.org/10.1016/j.ijimpeng.2021.104025

  4. W. Schonberg and S. Ryan, “Predicting metallic armour performance when impacted by fragment-simulating projectiles – Model adjustments and improvements,” Int. J. Impact Eng. 161, 104090 (2022). https://doi.org/10.1016/j.ijimpeng.2021.104090

  5. C. E. Anderson, Jr., “Analytical models for penetration mechanics: A review,” Int. J. Impact Eng. 108, 3–26 (2017). https://doi.org/10.1016/j.ijimpeng.2017.03.018

    Article  Google Scholar 

  6. K. D. Dhote and P. N. Verma, “Investigation of hole formation by steel sphere impacting on thin plate at hypervelocity,” Thin-Walled Struct. 126, 38–47 (2018). https://doi.org/10.1016/j.tws.2017.05.012

    Article  Google Scholar 

  7. R. F. Recht and T. W. Ipson, “Ballistic perforation dynamics.” J. Appl. Mec. 30 (3), 384–390 (1963). https://doi.org/10.1115/1.3636566

    Article  ADS  Google Scholar 

  8. A. Tate, “A theory for the deceleration of long rods after impact,” J. Mech. Phys. Solids 15 (6), 387–399 (1967). https://doi.org/10.1016/0022-5096(67)90010-5

    Article  ADS  Google Scholar 

  9. A. Tate, “Long rod penetration models—Part II. Extensions to the hydrodynamic theory of penetration,” Int. J. Mech. Sci. 28 (9), 599–612 (1986). https://doi.org/10.1016/0020-7403(86)90075-5

    Article  Google Scholar 

  10. H. M. Wen and W. H. Sun, “Transition of plugging failure modes for ductile metal plates under impact by flat-nosed projectiles,” Mech. Based Des. Struct. Mach. 38 (1), 86–104 (2010). https://doi.org/10.1080/15397730903415892

    Article  Google Scholar 

  11. S. E. Jones and W. K. Rule, “On the optimal nose geometry for a rigid penetrator, including the effects of pressure-dependent friction,” Int. J. Impact Eng. 24 (4), 403–415 (2000). https://doi.org/10.1016/S0734-743X(99)00157-8

    Article  Google Scholar 

  12. X. W. Chen and Q. M. Li, “Deep penetration of a non-deformable projectile with different geometrical characteristics,” Int. J. Impact Eng. 27 (6), 619–637 (2002). https://doi.org/10.1016/S0734-743X(02)00005-2

    Article  Google Scholar 

  13. W. J. Jiao and X. W. Chen, “Influence of the mushroomed projectile’s head geometry on long-rod penetration,” Int. J. Impact Eng.” 148, 103769 (2021). https://doi.org/10.1016/j.ijimpeng.2020.103769

  14. X. H. Dai, K. H. Wang, M. R. Li, et al., “Rigid elliptical cross-section ogive-nose projectiles penetration into concrete targets,” Def. Technol. 17 (3), 800–811 (2021). https://doi.org/10.1016/j.dt.2020.05.011

    Article  Google Scholar 

  15. H. Y. Wei, X. F. Zhang, C. Liu, et al., “Oblique penetration of ogive-nosed projectile into aluminum alloy targets,” Int. J. Impact Eng. 148, 103745 (2021). https://doi.org/10.1016/j.ijimpeng.2020.103745

  16. H. Dong, X. X. Zhang, H. J. Wu, et al., “Trajectory prediction of the tapered projectile and grooved projectile in deep penetration,” Int. J. Impact Eng. 170, 104337 (2022). https://doi.org/10.1016/j.ijimpeng.2022.104337

  17. X. Q. Ma, Q. M. Zhang, and X. W. Zhang, “A model for rigid asymmetric ellipsoidal projectiles penetrating into metal plates,” Int. J. Impact Eng. 163, 104140 (2022). https://doi.org/10.1016/j.ijimpeng.2021.104140

  18. H. Y. Wei, X. F. Zhang, C. Liu, et al., “A three-dimensional penetration trajectory model for ogive-nosed projectile into metal targets,” Int. J. Impact Eng. 157, 103998 (2021). https://doi.org/10.1016/j.ijimpeng.2021.103998

  19. H. J. Wu, X. M. Deng, H. Dong, et al., “Three-dimensional trajectory prediction and analysis of elliptical projectile,” Int. J. Impact Eng. 174, 104497 (2023). https://doi.org/10.1016/j.ijimpeng.2023.104497

  20. L. Liu, Y. R. Fan, W. Li, et al., “Cavity dynamics and drag force of high-speed penetration of rigid spheres into 10wt% gelatin,” Int. J. Impact Eng. 50, 68–75 (2012). https://doi.org/10.1016/j.ijimpeng.2012.06.004

    Article  Google Scholar 

  21. G. Seisson, D. Hébert, L. Hallo, et al., “Penetration and cratering experiments of graphite by 0.5-mm diameter steel spheres at various impact velocities,” Int. J. Impact Eng. 70, 14–20 (2014). https://doi.org/10.1016/j.ijimpeng.2014.03.004

    Article  Google Scholar 

  22. R.L. Martineau, M.B. Prime, and T. Duffey, “Penetration of HSLA-100 steel with tungsten carbide spheres at striking velocities between 0.8 and 2.5km/s,” Int. J. Impact Eng. 30 (5), 505–520 (2004). https://doi.org/10.1016/S0734-743X(03)00080-0

    Article  Google Scholar 

  23. H. Peng, S. Q. Wang, X. W. Hu, et al., “Experimental Investigation on failure behaviors of G50 ultra-high strength steel targets struck by tungsten alloy spherical fragments at high velocity,” Front. Mater. 7, (2021). https://doi.org/10.3389/fmats.2020.618771

  24. J. C. Cheng, S. Zhang, Q. Liu, et al., “Ballistic impact experiments and modeling on impact cratering, deformation and damage of 2024-T4 aluminum alloy,” Int. J. Mech. Sci. 224, 107312 (2022). https://doi.org/10.1016/j.ijmecsci.2022.107312

  25. J. C. Cheng, S. Zhang, Q. Liu, et al., “Multiple ballistic impacts on 2024-T4 aluminum alloy by spheres: Experiments and modelling,” J. Mater. Sci. Technol. 94, 164–174 (2021). https://doi.org/10.1016/j.jmst.2021.04.012

    Article  Google Scholar 

  26. K. Wen, X. W. Chen, and D. Di, “Modeling on the shock wave in spheres hypervelocity impact on flat plates,” Defence Technol. 15 (4), 457–466 (2019). https://doi.org/10.1016/j.dt.2019.01.006

    Article  Google Scholar 

  27. J. X. Wang and N. Zhou, “Experimental and numerical study of the ballistic performance of steel fibre-reinforced explosively welded targets impacted by a spherical fragment,” Compos. Part B: Eng. 75, 65–72 (2015). https://doi.org/10.1016/j.compositesb.2015.01.023

    Article  Google Scholar 

  28. Y. L. Bian, Q. Liu, Z. D. Feng, et al., “High-speed penetration dynamics of polycarbonate,” Int. J. Mech. Sci. 223, 107250 (2022). https://doi.org/10.1016/j.ijmecsci.2022.107250

  29. Y. G. Gao, S. S. Feng, G. Y. Huang, et al., “Experimental study of the oblique impact and ricochet characteristics of cylindrical fragments,” Int. J. Impact Eng. 170, 104334 (2022). https://doi.org/10.1016/j.ijimpeng.2022.104334

  30. V. L. Fomin, A. I. Gulidov, G. A. Sapozshnikov, et al., High-Speed Interaction of Bodies (SO RAN, Novosibirsk,1999) [in Russian].

    Google Scholar 

  31. A. M. Bragov, V. V. Balandin, L. A. Igumnov, et al., “Solution of the problem on the expansion of a spherical cavity in terms of estimation of the resistance to the penetration of a solid into the soil,” Mech. Solids 57, 543–552 (2022). https://doi.org/10.3103/S0025654422030074

    Article  ADS  Google Scholar 

  32. V. G. Bazhenov, V. V. Balandin, S. S. Grigoryan, and V. L. Kotov, “Analysis of models for calculating the motion of solids of revolution of minimum resistance in soil media,” J. Appl. Math. Mech. 78 (1), 65–76 (2014). https://doi.org/10.1016/j.jappmathmech.2014.05.008

    Article  MathSciNet  Google Scholar 

  33. K. Yu. Osipenko, “Stability of spatial motion of a body of revolution in an elastoplastic,” Medium. Mech. Solids 47 (2), 212–220 (2012). https://doi.org/10.3103/S0025654412020082

    Article  ADS  Google Scholar 

  34. G. E. Yakunina, “Characteristic features of the high-velocity motion of bodies in dense media,” J. Appl. Math. Mech. 76 (3), 310–323 (2012). https://doi.org/10.1016/j.jappmathmech.2012.07.007

    Article  Google Scholar 

  35. M. J. Forrestal, K. Okajima, and V. K. Luk, “Penetration of 6061-T651 aluminum targets with rigid long rods.” J. Appl. Mech. 55, 755–760 (1988). https://doi.org/10.1115/1.3173718

    Article  ADS  Google Scholar 

  36. M. J. Forrestal, D.Y. Tzou, E. Askari, et al., “Penetration into ductile metal targets with rigid spherical-nose rods,” Int. J. Impact Eng. 16, 699–710 (1995). https://doi.org/10.1016/0734-743X(95)00005-U

    Article  Google Scholar 

  37. T. L. Warren and K. L. Poormon, “Penetration of 6061-T6511 aluminum targets by ogive-nosed VAR 4340 steel projectiles at oblique angles: experiments and simulations,” Int. J. Impact Eng. 25 (10), 993–1022 (2001). https://doi.org/10.1016/S0734-743X(01)00024-0

    Article  Google Scholar 

  38. D. Z. Yankelevsky, V. R. Feldgun, and Y. S. Karinski, “The optimal nose shape of a rigid projectile deeply penetrating into a solid target considering friction,” Int. J. Impact Eng. 162, 104157 (2022). https://doi.org/10.1016/j.ijimpeng.2022.104157

  39. R. Hill, The Mathematical Theory of Plasticity (Oxford, London, 1950).

    Google Scholar 

  40. T. L.Warren, S. J. Hanchak, and K. L. Poormon, “Penetration of limestone targets by ogive-nosed VAR 4340 steel projectiles at oblique angles: experiments and simulations,” Int. J. Impact Eng. 30 (10), 1307–1331 (2004). https://doi.org/10.1016/j.ijimpeng.2003.09.047

    Article  Google Scholar 

  41. E. C. Cho, “The volume of a tetrahedron,” Appl. Math. Lett. 8 (2), 71–73 (1995). https://doi.org/10.1016/0893-9659(95)00014-H

    Article  MathSciNet  Google Scholar 

  42. G. Yossifon, M. B. Rubin, and A. L. Yarin, “Penetration of a rigid projectile into a finite thickness elastic–plastic target-comparison between theory and numerical computations,” Int. J. Impact Eng. 25 (3), 265–290 (2001). https://doi.org/10.1016/S0734-743X(00)00040-3

    Article  Google Scholar 

  43. D. Q. Hu, R. Q. Chi, Y. Y. Liu, et al., “Sensitivity analysis of spacecraft in micrometeoroids and orbital debris environment based on panel method,” Def. Technol. 19 (1), 17 (2023). https://doi.org/10.1016/j.dt.2021.11.001

    Article  Google Scholar 

  44. T. L. Warren, “The effect of target inertia on the penetration of aluminum targets by rigid ogive-nosed long rods,” Int. J. Impact Eng. 91 (5), 6–13 (2016). https://doi.org/10.1016/j.ijimpeng.2015.12.007

    Article  Google Scholar 

Download references

ACKNOWLEDGMENTS

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper. The author wants to thank Zhang P for reviewing this article and making suggestions to improve it.

Funding

No funding was received for conducting this study.

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to T. L. Liu or Y. X. Xu.

Ethics declarations

The authors declare that they have no conflicts of interest.

Additional information

Publisher’s Note.

Allerton Press remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

APPENDIX

APPENDIX

The resistance decay model in this paper is derived based on the Forrestal’s spherical dynamic cavity expansion model. The derivation process [44] of Forrestal cavity expansion model is introduced before the derivation of the resistance decay model.

If r is the Euler spherical coordinates, u is the outward positive radial displacement of the particle, the radius of the cavity is \(r - u\). Then, the equilibrium equation of axisymmetric sphere during expansion is:

$$\frac{{\partial {{\sigma }_{{rr}}}}}{{\partial r}} + \frac{2}{r}({{\sigma }_{{rr}}} - {{\sigma }_{{\theta \theta }}}) = {{\rho }_{t}}\frac{{d{v}}}{{dt}} = - {{\rho }_{t}}\left( {\frac{{\partial {v}}}{{\partial t}} + {v}\frac{{\partial {v}}}{{\partial r}}} \right).$$
(A.1)

Stresses are assumed to be positive in compression, the continuity equation is:

$$\frac{{\partial {v}}}{{\partial r}} + 2\frac{{v}}{r} = 0.$$
(A.2)

When the cavity expands, the target plate material satisfies the mass conservation equation so:

$${{(r - u)}^{3}} = {{r}^{3}} - {{a}^{3}}.$$
(A.3)

The radial displacement is:

$$u = r\left[ {1 - {{{\left( {1 - \frac{{{{a}^{3}}}}{{{{r}^{3}}}}} \right)}}^{{1/3}}}} \right].$$
(A.4)

In the instantaneous state, the strain of the cavity wall is ignored, as follows:

$$\frac{{\partial u}}{{\partial t}} = \frac{{du}}{{dt}} = \frac{r}{3}{{\left( {1 - \frac{{{{a}^{3}}}}{{{{r}^{3}}}}} \right)}^{{ - \frac{2}{3}}}}\left( { - 3\frac{{{{a}^{2}}\dot {a}}}{{{{r}^{3}}}}} \right) = {{\left( {1 - \frac{{{{a}^{3}}}}{{{{r}^{3}}}}} \right)}^{{ - \frac{2}{3}}}}\frac{{{{a}^{2}}\dot {a}}}{{{{r}^{2}}}}$$
(A.5)

The equation of strain is:

$$\frac{{\partial u}}{{\partial r}} = - 2{{\left( {1 - \frac{{{{a}^{3}}}}{{{{r}^{3}}}}} \right)}^{{ - \frac{2}{3}}}}\frac{{{{a}^{2}}\dot {a}}}{{{{r}^{3}}}} + \frac{{{{a}^{2}}\dot {a}}}{{{{r}^{2}}}}\left( {2{{{\left( {1 - \frac{{{{a}^{3}}}}{{{{r}^{3}}}}} \right)}}^{{ - \frac{5}{3}}}}\frac{{{{a}^{3}}}}{{{{r}^{4}}}}} \right).$$
(A.6)

Ignoring small deformation, radial strain rate can be written as:

$${{\dot {\varepsilon }}_{{rr}}} = - \frac{{\partial {v}}}{{\partial r}} = \frac{{2{{a}^{2}}\dot {a}}}{{{{r}^{3}}}}.$$
(A.7)

And the circumferential and tangential strain rates are:

$${{\dot {\varepsilon }}_{{\theta \theta }}} = {{\dot {\varepsilon }}_{{\varphi \varphi }}} = - \frac{{v}}{r}.$$
(A.8)

The radial engineering strain is:

$${{\varepsilon }_{{rr}}}{\text{ = }}\ln \left( {1 - \frac{{du}}{{dr}}} \right)$$
(A.9)

Tangential engineering strain is:

$${{\varepsilon }_{{\theta \theta }}} = \ln \left( {1 - \frac{u}{r}} \right).$$
(A.10)

Considering that the constitutive model of target plate material is power hardening model, we can get:

$$\sigma = \left\{ \begin{gathered} E\varepsilon ,\quad \sigma < Y \hfill \\ Y{{\left( {\frac{{E\varepsilon }}{Y}} \right)}^{n}},\quad \sigma \geqslant Y \hfill \\ \end{gathered} \right..$$
(A.11)

By substituting Eqs. (A.5)–(A.11) into equilibrium Eq. (A.1), we can get:

$$\frac{{\partial {{\sigma }_{{rr}}}}}{{\partial r}} = - \frac{{2Y}}{r}{{\left( {\frac{{2E}}{{3Y}}} \right)}^{n}}{{\left[ { - \ln \left( {1 - {{{\left( {\frac{a}{r}} \right)}}^{3}}} \right)} \right]}^{n}} - {{\rho }_{t}}\left( {\frac{{\ddot {a}{{a}^{2}} + 2a{{{\dot {a}}}^{2}}}}{{{{r}^{2}}}} - \frac{{2{{a}^{4}}{{{\dot {a}}}^{2}}}}{{{{r}^{5}}}}} \right).$$
(A.12)

By integrating the above Eq. (A.12), we can get:

$${{\sigma }_{{rr}}}(a) = \int\limits_a^c {\frac{{2Y}}{r}{{{\left( {\frac{{2E}}{{3Y}}} \right)}}^{n}}{{{\left[ { - \ln \left( {1 - {{{\left( {\frac{a}{r}} \right)}}^{3}}} \right)} \right]}}^{n}} - {{\rho }_{t}}\left( {\frac{{\ddot {a}{{a}^{2}} + 2a{{{\dot {a}}}^{2}}}}{{{{r}^{2}}}} - \frac{{2{{a}^{4}}{{{\dot {a}}}^{2}}}}{{{{r}^{5}}}}} \right)dr} .$$
(A.13)

Complete integration of the above equation can be obtained:

$${{\sigma }_{{rr}}}(a) = {{\sigma }_{{rr}}}(c) + 2Y{{\left( {\frac{{2E}}{{3Y}}} \right)}^{n}}\int\limits_a^c {{{{\left[ { - \ln \left( {1 - {{{\left( {\frac{a}{r}} \right)}}^{3}}} \right)} \right]}}^{n}}\frac{{dr}}{r} + {{\rho }_{r}}\left[ { - (\ddot {a}{{a}^{2}} + 3a{{{\dot {a}}}^{2}})\left( {\frac{1}{c} - \frac{1}{a}} \right) + \frac{{{{a}^{4}}{{{\dot {a}}}^{2}}}}{2}\left( {\frac{1}{{{{c}^{4}}}} - \frac{1}{{{{a}^{4}}}}} \right)} \right]} ,$$
(A.14)

where \({{\sigma }_{{rr}}}(c)\) is the elastic-plastic boundary pressure.

At first, we solve for the stress in the elastic region. According to the elastic equilibrium equation. Eq. (A.3) can be obtained as follows:

$${{r}^{3}} - 3{{r}^{2}}u + 3r{{u}^{2}} - {{u}^{3}} = {{r}^{3}} - {{a}^{3}}.$$
(A.15)

Since the elastic region displacement u is very small, the quadratic term u in Eq. (A.15) is removed, and the deformation can be obtained as follows:

$$u = \frac{{{{a}^{3}}}}{{3{{r}^{2}}}}.$$
(A.16)

Then the radial strain at the elastic region is:

$${{\varepsilon }_{{rr}}} = \frac{{du}}{{dr}} = \frac{{2{{a}^{3}}}}{{3{{r}^{3}}}}.$$
(A.17)

The circumferential strain in the elastic region is:

$${{\varepsilon }_{{\theta \theta }}} = {{\varepsilon }_{{\varphi \varphi }}} = \frac{{du}}{{dr}} = \frac{{2{{a}^{3}}}}{{3{{r}^{3}}}}.$$
(A.18)

According to generalized Hooke’s law of elastic stage:

$${{\sigma }_{{rr}}} - {{\sigma }_{{\theta \theta }}} = 2G({{\varepsilon }_{{rr}}} - {{\varepsilon }_{{\theta \theta }}}).$$
(A.19)

By substituting Eq. (A.18) into Eq. (A.19), we can get:

$${{\sigma }_{{rr}}} - {{\sigma }_{{\theta \theta }}} = 2G\left( {\frac{{2{{a}^{3}}}}{{3{{r}^{3}}}} - \left( { - \frac{{{{a}^{3}}}}{{3{{r}^{3}}}}} \right)} \right) = 2G\left( {\frac{{{{a}^{3}}}}{{{{r}^{3}}}}} \right).$$
(A.20)

The relationship between shear modulus G and elastic modulus E is as follows:

$$G = \frac{E}{{2(1 + \nu )}}.$$
(A.21)

For metal materials, Poisson’s ratio is 0.5. Then, Eqs. (A.16)–(A.21) are substituted into the stress balance Eq. (A.1) to obtain:

$$\frac{{d{{\sigma }_{{rr}}}}}{{dr}} = - \frac{2}{r}({{\sigma }_{{rr}}} - {{\sigma }_{{\theta \theta }}}) - {{\rho }_{t}}\left( {\frac{{\partial u}}{{\partial t}} + {v}\frac{{\partial u}}{{\partial r}}} \right) = - \frac{{4E{{a}^{3}}}}{{3{{r}^{2}}}} - {{\rho }_{t}}\left( {\frac{{\ddot {a}{{a}^{2}} + 2a{{{\dot {a}}}^{2}}}}{{{{r}^{2}}}} - \frac{{2{{a}^{4}}{{{\dot {a}}}^{2}}}}{{{{r}^{5}}}}} \right).$$
(A.22)

Integral:

$${{\sigma }_{{rr}}}(c) = \frac{{4E{{a}^{3}}}}{{9{{c}^{3}}}} + {{\rho }_{t}}(\ddot {a}{{a}^{2}} + 2a{{\dot {a}}^{2}})\left( {\frac{1}{k} - \frac{1}{c}} \right) - \frac{{{{a}^{4}}{{{\dot {a}}}^{2}}}}{2}\left( {\frac{1}{{{{k}^{4}}}} - \frac{1}{{{{c}^{4}}}}} \right).$$
(A.23)

According to the position relation of the elastic-plastic region shown in Fig. 4, for the case of penetration of the semi-infinite target, there is: \(k \to \infty \). Therefore, the dynamic resistance term of the elastic region is:

$${{\rho }_{t}}(\ddot {a}{{a}^{2}} + 2a{{\dot {a}}^{2}})\left( { - \frac{1}{c}} \right) + \frac{{{{a}^{4}}{{{\dot {a}}}^{2}}}}{2}\left( {\frac{1}{{{{c}^{4}}}}} \right).$$
(A.24)

Meanwhile, according to Eq. (A.14), the dynamic resistance term of the plastic region is:

$$ - {{\rho }_{t}}(\ddot {a}{{a}^{2}} + 2a{{\dot {a}}^{2}})\left( {\frac{1}{c} - \frac{1}{a}} \right) - \frac{{{{a}^{4}}{{{\dot {a}}}^{2}}}}{2}\left( {\frac{1}{{{{c}^{4}}}} - \frac{1}{{{{a}^{4}}}}} \right).$$
(A.25)

The sum of the above two formulas is the dynamic resistance term of the dynamic cavity expansion:

$${{\rho }_{t}}(\ddot {a}{{a}^{2}} + 2a{{\dot {a}}^{2}})\left( {\frac{1}{k} - \frac{1}{a}} \right) - \frac{{{{a}^{4}}{{{\dot {a}}}^{2}}}}{2}\left( {\frac{1}{{{{k}^{4}}}} - \frac{1}{{{{a}^{4}}}}} \right).$$
(A.26)

According to Eq. (A.14), the static resistance term of the plastic region is:

$$2Y{{\left( {\frac{{2E}}{{3Y}}} \right)}^{n}}\int\limits_a^c {{{{\left[ { - \ln \left( {1 - {{{\left( {\frac{a}{r}} \right)}}^{3}}} \right)} \right]}}^{n}}\frac{{dr}}{r}} .$$
(A.27)

According to Eqs. (A.23) and (A.27), the sum of static resistance terms of the elastic-plastic interface and the plastic region is:

$${{\sigma }_{{rr}}}(a) = \frac{{4E{{a}^{3}}}}{{9{{c}^{3}}}} + 2Y{{\left( {\frac{{2E}}{{3Y}}} \right)}^{n}}\int\limits_a^c {\frac{{{{{\left[ { - \ln {{{\left( {1 - \frac{a}{r}} \right)}}^{3}}} \right]}}^{n}}}}{r}dr} .$$
(A.28)

Let \(x = 1 - {{\left( {\frac{a}{r}} \right)}^{3}}\), substitute into Eq. (A.28) to get:

$${{\sigma }_{{rr}}}(a) = \frac{{4E{{a}^{3}}}}{{9{{c}^{3}}}} + \frac{{2Y}}{3}{{\left( {\frac{{2E}}{{3Y}}} \right)}^{n}}\int\limits_0^{1 - {{{\left( {\frac{a}{c}} \right)}}^{3}}} {\frac{{{{{[ - \ln x]}}^{n}}}}{{1 - x}}} dx.$$
(A.29)

If the moving velocity of the cavity wall is \({{{v}}_{c}}\), and the displacement of the cavity wall is: \(a = {{{v}}_{c}}t\), then the displacement of the elastic-plastic interface is:

$$c = {{\left( {\frac{{2E}}{{3Y}}} \right)}^{{1/3}}}{{{v}}_{c}}t.$$
(A.30)

We can get:

$$\frac{a}{c} = {{\left( {\frac{{3Y}}{{2E}}} \right)}^{{1/3}}}.$$
(A.31)

Then the static resistance of the projectile and target contact surface is:

$${{\sigma }_{{rr}}}(a) = \frac{{2Y}}{3} + \frac{{2Y}}{3}{{\left( {\frac{{2E}}{{3Y}}} \right)}^{n}}\int\limits_0^{1 - {{{\left( {\frac{a}{c}} \right)}}^{3}}} {\frac{{{{{[ - \ln x]}}^{n}}}}{{1 - x}}} dx.$$
(A.32)

Equation (A.32) is hill’s static cavity expansion resistance model, corresponding to Eq. (2.6) in paper. Eq. (A.32) and Eq. (A.26) are superimposed and summed to obtain the dynamic cavity expansion resistance model:

$${{\sigma }_{{rr}}}(a) = \frac{{2Y}}{3} + \frac{{2Y}}{3}{{\left( {\frac{{2E}}{{3Y}}} \right)}^{n}}\int\limits_0^{1 - {{{\left( {\frac{a}{c}} \right)}}^{3}}} {\frac{{{{{[ - \ln x]}}^{n}}}}{{1 - x}}} dx + {{\rho }_{t}}\left[ {(\ddot {a}{{a}^{2}} + 2a{{{\dot {a}}}^{2}})\left( {\frac{1}{k} - \frac{1}{a}} \right) - \frac{{{{a}^{4}}{{{\dot {a}}}^{2}}}}{2}\left( {\frac{1}{{{{k}^{4}}}} - \frac{1}{{{{a}^{4}}}}} \right)} \right].$$
(A.33)

When the semi-infinite target is penetrated, \(k \to \infty \), \(\frac{1}{k}\) and \(\frac{1}{{{{k}^{4}}}}\) is zero, then the above equation is simplified as:

$${{\sigma }_{{rr}}}(a) = \frac{{2Y}}{3} + \frac{{2Y}}{3}{{\left( {\frac{{2E}}{{3Y}}} \right)}^{n}}\int\limits_0^{1 - {{{\left( {\frac{a}{c}} \right)}}^{3}}} {\frac{{{{{[ - \ln x]}}^{n}}}}{{1 - x}}} dx + {{\rho }_{t}}\left( {\ddot {a}a + \frac{3}{2}{{{\dot {a}}}^{2}}} \right).$$
(A.34)

Acceleration term \(\ddot {a}\) and displacement term a are included in the above equation. Since we cannot measure the acceleration term \(\ddot {a}\), the acceleration term \(\ddot {a}\) is removed in practical work, and then the model becomes:

$${{\sigma }_{{rr}}}(a) = \frac{{2Y}}{3}\left[ {1 + {{{\left( {\frac{{2E}}{{3Y}}} \right)}}^{n}}\int\limits_0^{1 - \frac{{3Y}}{{2E}}} {\frac{{{{{[ - \ln x]}}^{n}}}}{{1 - x}}} dx} \right] + \beta {{\rho }_{t}}{{\dot {a}}^{2}}.$$
(A.35)

When \(\beta \) takes as 1.5, Eq. (A.35) is the Forrestal resistance model, see Eq. (2.7).

The Warren resistance decay model [33] is based on the position relation between the d distance from any point on the surface of the projectile to the expansion elastic-plastic boundary of the target cavity and b the distance to the free boundary of the target plate. We adopt the same judgment method in this paper, but the difference is that in Warren’s paper [33], the target plate material constitutive model is an ideal plastic material model, which satisfies \({{\sigma }_{{rr}}} - {{\sigma }_{{\theta \theta }}} = Y\). In order to ensure the consistency of the resistance model and the resistance decay model, the power hardening model is adopted in this paper, as shown in Eq. (A.11).

When the distance between the projectile surface and the elastic-plastic interface of the target plate is greater than the distance between the projectile surface and the free boundary of the target plate, i.e. \(d > b\), according to Eq. (A.33), the acceleration term \(\ddot {a}\) is ignored, thus:

$$\sigma (d) = \frac{{2Y}}{3}\left( {1 - {{{\left( {\frac{c}{d}} \right)}}^{3}}} \right) + \frac{{2Y}}{3}{{\left( {\frac{{2E}}{{3Y}}} \right)}^{n}}\int\limits_0^{1 - \frac{{3Y}}{{2E}}} {\frac{{{{{[ - \ln x]}}^{n}}}}{{1 - x}}dx} ] + {{\rho }_{t}}{{\dot {a}}^{2}}\left[ {\frac{3}{2} - \frac{{2a}}{d} + \frac{1}{2}\frac{{{{a}^{4}}}}{{{{d}^{4}}}}} \right].$$
(A.36)

In this case, the elastoplastic interface has not reached the free boundary of the target plate, and the stress in the plastic deformation region has not been affected. When the distance between the projectile surface and the elastic-plastic interface of the target plate is less than the distance between the projectile surface and the free boundary of the target plate, i.e. \(d < b\). According to Eq. (A.33), ignoring the acceleration term \(\ddot {a}\), we can get:

$$\sigma (d) = \frac{{2Y}}{3}{{\left( {\frac{{2E}}{{3Y}}} \right)}^{n}}\int\limits_0^{1 - {{{\left( {\frac{a}{d}} \right)}}^{3}}} {\frac{{{{{[ - \ln x]}}^{n}}}}{{1 - x}}dx} ] + {{\rho }_{t}}{{\dot {a}}^{2}}\left[ {\frac{3}{2} - \frac{{2a}}{d} + \frac{1}{2}\frac{{{{a}^{4}}}}{{{{d}^{4}}}}} \right].$$
(A.37)

In this case, the stress in the elastic deformation region is 0, and the stress in the plastic deformation region is affected.

For the semi-infinite target penetration problem, the Forrestal model is applied to the resistance model, as follows:

$$\sigma = \frac{{2Y}}{3}\left[ {1 + {{{\left( {\frac{{2E}}{{3Y}}} \right)}}^{n}}\int\limits_0^{1 - \frac{{3Y}}{{2E}}} {\frac{{{{{[ - \ln x]}}^{n}}}}{{1 - x}}dx} } \right] + \frac{3}{2}{{\rho }_{t}}{{\dot {a}}^{2}}.$$
(A.38)

Therefore, the resistance decay coefficient can be written as: \(\sigma (d){\text{/}}\sigma \). Written as a piecewise function:

$$\left\{ \begin{gathered} f(d,a,\ddot {a}) = \left\{ \begin{gathered} A{\text{/}}C\begin{array}{*{20}{c}} {}&{}&{d > b} \end{array} \hfill \\ B{\text{/}}C\begin{array}{*{20}{c}} {}&{}&{d < b} \end{array} \hfill \\ \end{gathered} \right. \hfill \\ A = \frac{{2Y}}{3}\left( {1 - {{{\left( {\frac{c}{d}} \right)}}^{3}}} \right) + \frac{{2Y}}{3}{{\left( {\frac{{2E}}{{3Y}}} \right)}^{n}}\int\limits_0^{1 - \frac{{3Y}}{{2E}}} {\frac{{{{{[ - \ln x]}}^{n}}}}{{1 - x}}dx} ] + {{\rho }_{t}}{{{\dot {a}}}^{2}}\left[ {\frac{3}{2} - \frac{{2a}}{d} + \frac{1}{2}\frac{{{{a}^{4}}}}{{{{d}^{4}}}}} \right] \hfill \\ B = \frac{{2Y}}{3}{{\left( {\frac{{2E}}{{3Y}}} \right)}^{n}}\int\limits_0^{1 - {{{\left( {\frac{a}{d}} \right)}}^{3}}} {\frac{{{{{[ - \ln x]}}^{n}}}}{{1 - x}}dx} ] + {{\rho }_{t}}{{{\dot {a}}}^{2}}\left[ {\frac{3}{2} - \frac{{2a}}{d} + \frac{1}{2}\frac{{{{a}^{4}}}}{{{{d}^{4}}}}} \right] \hfill \\ C = \frac{{2Y}}{3}\left[ {1 + {{{\left( {\frac{{2E}}{{3Y}}} \right)}}^{n}}\int\limits_0^{1 - \frac{{3Y}}{{2E}}} {\frac{{{{{[ - \ln x]}}^{n}}}}{{1 - x}}dx} } \right] + \frac{3}{2}{{\rho }_{t}}{{{\dot {a}}}^{2}} \hfill \\ \end{gathered} \right..$$
(A.39)

The above equation is Eqs. (2.9)(2.12) in Section 2.4. Derivation is completed

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Liu, T.L., Xu, Y.X. & Wang, Y.F. Oblique Penetration of Spherical Projectile into Low-Carbon Steel Target: Experiment, Theory, 3D Penetration Model. Mech. Solids 58, 2295–2318 (2023). https://doi.org/10.3103/S0025654423601039

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.3103/S0025654423601039

Keywords:

Navigation