Skip to main content

Advertisement

Log in

Spall Characterization in Epoxy Via Laser Spallation

  • Research paper
  • Published:
Experimental Mechanics Aims and scope Submit manuscript

Abstract

Background: The strength of materials under extreme dynamic loading conditions, such as in the case of shock wave loading, is assessed from their spallation characteristics. Under laboratory conditions, flyer plate impact, or sometimes laser-induced stress waves, is employed to instigate spall in a material. These methods are often combined with velocity interferometer system for any reflector (VISAR) technique for performing transient measurements. Although the VISAR can record the velocity of extremely fast-moving surfaces, it requires a complex optical setup and a specialized data reduction technique. Objective: In this study, a simpler approach is adopted by extending laser spallation method to determine the spall strength of epoxy, while performing in situ interferometric measurements, directly on top of thick epoxy films. Methods: The glass/epoxy test samples are prepared by transferring an aluminum coating on top of epoxy layers with different thicknesses. Laser-induced stress waves transmit across the substrate/film interface and induce subsurface failure in the epoxy at sufficiently high incident laser energy. The nature and magnitude of the waves are deciphered from the out-of-plane displacement histories of the top reflective sample surfaces, which are recorded by using a Michelson interferometer. Results: The interferometric data reveal the development of two (temporally) well-separated stress waves: an ablation-induced high-amplitude short-duration longitudinal pulse, which is referred to as the primary wave, and a secondary wave, which travels at a comparatively slower speed. The complex constructive interaction of the two waves develops a high-magnitude tensile stress region in the epoxy layer. The spall strength is quantified by superimposing the two stress wave histories associated with the critical energy fluence. Conclusions: The spall depths predicted from spatiotemporal wave travel analyses are in excellent agreement with the experimental observations. The newly adopted methodology estimates the spall strength of epoxy as 260 ± 20 MPa.

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

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Institutional subscriptions

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

Similar content being viewed by others

Notes

  1. Lev and Argon [35] suggested that the wave front remains nearly planar up to a distance of about 2 to 3 times the radius of the region in which the energy is impinged. For samples that are less than 1.75-mm-thick, when a 2-mm-diameter Nd:YAG energy spot is used, a planar wave front can be safely assumed.

  2. In a separate study, SHPB experiments were performed to evaluate the dynamic compression response of rigid particle-filled polymers [36]. The yield strength (σY) of the epoxy at a strain rate of 1550/s was evaluated to be ~220 MPa. By employing the equation, \( {\sigma}_{HEL}=\frac{\left(1-\nu \right)}{\left(1-2\nu \right)}{\sigma}_Y \) (as reported in [37]), the Hugoniot elastic limit (σHEL) of epoxy is calculated to be ~500 MPa.

  3. The average longitudinal wave speed in epoxy obtained from interferometric measurements (~2675 m/s) is marginally higher than the wave speeds obtained from ultrasonic measurements, which are discussed in Sect. ‘Wave Speeds in the Bulk Material’ (~2610 m/s).

References

  1. Eskandari H, Nemes JA (2000) Dynamic testing of composite laminates with a tensile split Hopkinson bar. J Compos Mater 34(4):260–273

    Article  Google Scholar 

  2. Dıaz-Rubio FG, Perez JR, Galvez VS (2002) The spalling of long bars as a reliable method of measuring the dynamic tensile strength of ceramics. Int J Impact Eng 27(2):161–177

    Article  Google Scholar 

  3. Gilat A, Goldberg RK, Roberts GD (2002) Experimental study of strain-rate-dependent behavior of carbon/epoxy composite. Compos Sci Technol 62(10–11):1469–1476

    Article  Google Scholar 

  4. Wu H, Zhang Q, Huang F, Jin Q (2005) Experimental and numerical investigation on the dynamic tensile strength of concrete. Int J Impact Eng 32(1–4):605–617

    Article  Google Scholar 

  5. Schuler H, Mayrhofer C, Thoma K (2006) Spall experiments for the measurement of the tensile strength and fracture energy of concrete at high strain rates. Int J Impact Eng 32(10):1635–1650

    Article  Google Scholar 

  6. McQueen RG, Marsh SP (1962) Ultimate yield strength of copper. J Appl Phys 33(2):654–665

    Article  Google Scholar 

  7. Kanel GI, Rasorenov SV, Fortov VE (1991) The failure waves and spallations in homogeneous brittle materials. In: Schmidt SC et al (eds) Shock compression of condensed matter. Elsevier, New York, pp 451–454

    Google Scholar 

  8. Grady DE (1998) Shock-wave compression of brittle solids. Mech Mater 29(3–4):181–203

    Article  Google Scholar 

  9. Yuan F, Tsai L, Prakash V, Rajendran AM, Dandekar DP (2007) Spall strength of glass fiber reinforced polymer composites. Int J Solids Struct 44(24):7731–7747

    Article  Google Scholar 

  10. Katz S, Zaretsky E, Grossman E, Wagner HD (2009) Dynamic tensile strength of organic fiber-reinforced epoxy micro-composites. Compos Sci Technol 69(7–8):1250–1255

    Article  Google Scholar 

  11. Bie BX, Han JH, Lu L, Zhou XM, Qi ML, Zhang Z, Luo SN (2015) Dynamic fracture of carbon nanotube/epoxy composites under high strain-rate loading. Compos A: Appl Sci Manuf 68:282–288

    Article  Google Scholar 

  12. Xie W, Zhang W, Guo L, Gao Y, Li D, Jiang X (2018) The shock and spallation behavior of a carbon fiber reinforced polymer composite. Compos Part B 153:176–183

    Article  Google Scholar 

  13. Rosenberg Z, Luttwak G, Yeshurun Y, Partom Y (1983) Spall studies of differently treated 2024 A1 specimens. J Appl Phys 54(5):2147–2152

    Article  Google Scholar 

  14. Millett JCF, Bourne NK (2006) Shock and release of polycarbonate under one-dimensional strain. J Mater Sci 41(6):1683–1690

    Article  Google Scholar 

  15. Eliezer S, Gilath I, Bar-Noy T (1990) Laser-induced spall in metals: experiment and simulation. J Appl Phys 67(2):715–724

    Article  Google Scholar 

  16. Fortov VE, Kostin VV, Eliezer S (1991) Spallation of metals under laser irradiation. J Appl Phys 70(8):4524–4531

    Article  Google Scholar 

  17. Tollier L, Fabbro R, Bartnicki E (1998) Study of the laser-driven spallation process by the velocity interferometer system for any reflector interferometry technique. I Laser-shock characterization. J Appl Phys 83(3):1224–1230

    Article  Google Scholar 

  18. Tollier L, Fabbro R (1998) Study of the laser-driven spallation process by the VISAR interferometry technique. II Experiment and simulation of the spallation process. J Appl Phys 83(3):1231–1237

    Article  Google Scholar 

  19. Gay E, Berthe L, Boustie M, Arrigoni M, Trombini M (2014) Study of the response of CFRP composite laminates to a laser-induced shock. Compos Part B 64:108–115

    Article  Google Scholar 

  20. Yang LC (1974) Stress waves generated in thin metallic films by a Q-switched ruby laser. J Appl Phys 45(6):2601–2608

    Article  Google Scholar 

  21. Vossen JL (1978) measurements of film-substrate bond strength by laser spallation. In adhesion measurement of thin films, thick films, and bulk coatings. ASTM International

  22. Gupta V, Argon AS, Parks DM, Cornie JA (1992) Measurement of interface strength by a laser spallation technique. J Mech Phys Solids 40(1):141–180

    Article  Google Scholar 

  23. Gupta V, Yuan J, Pronin A (1994) Recent developments in the laser spallation technique to measure the interface strength and its relationship to interface toughness with applications to metal/ceramic, ceramic/ceramic and ceramic/polymer interfaces. J Adhes Sci Technol 8(6):713–747

    Article  Google Scholar 

  24. Youssef G, Gupta V (2012) Dynamic tensile strength of polyurea. J Mater Res 27(2):494–499

    Article  Google Scholar 

  25. Wang J, Weaver RL, Sottos NR (2002) A parametric study of laser induced thin film spallation. Exp Mech 42(1):74–83

    Article  Google Scholar 

  26. Gupta V, Kireev V, Tian J, Yoshida H, Akahoshi H (2003) Glass-modified stress waves for adhesion measurement of ultra-thin films for device applications. J Mech Phys Solids 51(8):1395–1412

    Article  Google Scholar 

  27. Kandula SSV, Hartfield CD, Geubelle PH, Sottos NR (2008) Adhesion strength measurement of polymer dielectric interfaces using laser spallation technique. Thin Solid Films 516(21):7627–7635

    Article  Google Scholar 

  28. Kitey R, Sottos NR, Geubelle PH (2010) A hybrid experimental/numerical approach to characterize interfacial adhesion in multilayer low-κ thin film specimens. Thin Solid Films 519(1):337–344

    Article  Google Scholar 

  29. Zhou M, Zhang YK, Cai L (2002) Adhesion measurement of thin films by a modified laser spallation technique: theoretical analysis and experimental investigation. Appl Phys A 74(4):475–480

    Article  Google Scholar 

  30. Gupta V, Hernandez R, Charconnet P (2001) Effect of humidity and temperature on the tensile strength of polyimide/silicon nitride interface and its implications for electronic device reliability. Mater Sci Eng A 317(1–2):249–256

    Article  Google Scholar 

  31. Grady ME, Geubelle PH, Sottos NR (2014) Interfacial adhesion of photodefinable polyimide films on passivated silicon. Thin Solid Films 552:116–123

    Article  Google Scholar 

  32. Grady ME, Beiermann BA, Moore JS, Sottos NR (2014) Shockwave loading of mechanochemically active polymer coatings. ACS Appl Mater Interfaces 6(8):5350–5355

    Article  Google Scholar 

  33. Singh SS, Kitey R (2017) Effect of interface profile and incident wave characteristics on aluminum/epoxy dynamic adhesion strength. Int J Adhes Adhes 79:8–17

    Article  Google Scholar 

  34. Barker LM (1972) Laser interferometry in shock-wave research. Exp Mech 12(5):209–215

    Article  Google Scholar 

  35. Lev LC, Argon AS (1996) Spallation of thin elastic coatings from elastic substrates by laser induced pressure pulses. J Appl Phys 80(1):529–542

    Article  Google Scholar 

  36. Singh SS, Parameswaran V, Kitey R (2018) Dynamic compression behavior of glass filled epoxy composites: influence of filler shape and exposure to high temperature. Compos Part B 164:103–115

    Article  Google Scholar 

  37. Rosenberg Z (1993) On the relation between the Hugoniot elastic limit and the yield strength of brittle materials. J Appl Phys 74(1):752–753

    Article  Google Scholar 

  38. Asay JR, Shahinpoor M (eds) (2012) High-pressure shock compression of solids, Chapter-2. Springer Science & Business Media, New York

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to R. Kitey.

Ethics declarations

Conflict of Interest

The authors declare that they have no conflict of interest.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A: Laser-induced failure at the glass/epoxy interface

The development of primary and secondary waves in glass/epoxy samples upon the incidence of laser energy is discussed in Sect. ‘Interferometric Measurements’. To exclude the influence of secondary waves on the stress field, the experiments are performed on a sample prepared by depositing a 4-μm-thick epoxy layer on a glass substrate. This thin epoxy layer will experience at least five back-and-forth reflections of the longitudinal wave before the secondary wave appears at the glass/epoxy interface (in 8 ns). The failure initiation micrograph included in Fig. 14a corresponds to a YAG energy incidence of 65 mJ/mm2. The spallation image at 120 mJ/mm2 and its associated surface profile (illustrated in (b) and (c)) suggest glass/epoxy interfacial failure.

Fig. 14
figure 14

a The glass/epoxy interfacial failure initiation and b complete spallation, observed in a 4-μm-thick epoxy film. c Profilometric scan across the spalled region, shown in (b)

Representative fringes obtained by conducting experiments on a glass/Al calibration sample at 60 mJ/mm2 (just before failure initiation) are shown in Fig. 15a. The data are analyzed by following steps, as discussed in Sect. ‘Experimental Details’, and the longitudinal pulse history in the substrate is plotted in Fig. 15b. The interfacial stress history shown in Fig. 15c is obtained by performing computations on a sample with a 4-μm-thick epoxy layer. The modeling details are discussed in Sect. ‘Evaluating Stress Field in Epoxy’. The glass/epoxy interfacial strength is inferred to be ~200 MPa from the peak value of the interfacial stress.

Fig. 15
figure 15

a The representative interferometric data obtained by conducting laser spallation experiments on the glass/aluminum calibration sample at 60 mJ/mm2 laser fluence. b The substrate stress history associated to the primary wave. c Computationally obtained glass/epoxy interface stress history (with 4-μm-thick epoxy layer) corresponding to the transient loading shown in (b)

Stress history in epoxy prior to interfacial failure

The computationally obtained stress field in the epoxy layer is analyzed next to estimate the maximum tensile stress in the epoxy film prior to failure initiation at the substrate/film interface. The substrate stress corresponding to a laser energy of 50 mJ/mm2, as shown in Fig. 16a, is employed as the loading pulse in the 4-μm-thick epoxy layer model.

Fig. 16
figure 16

a The stress history in a glass substrate obtained by performing interferometric measurements at 50 mJ/mm2 laser fluence, b Computationally obtained stress field in a 4-μm-thick epoxy layer at various times. A zero ns refers to the time when the stress wave begins to transmit into the epoxy layer

The stress histories in epoxy at various times are illustrated in Fig. 16b. The labeled times are measured with respect to when the longitudinal pulse begins to transmit into the epoxy layer. The values of 0 and 4 μm on the x-axis correspond to the glass/epoxy interface and the free surface of the epoxy layer, respectively. After ~4 ns, a tensile region started to develop in the epoxy film. The figure clearly shows that by 4.75 ns, the entire epoxy layer is subjected to tensile stress. At 5.75 ns, the maximum tensile stress is noted to be ~175 MPa. Further continuing the simulations reveals a drop in the magnitude of the stress field (not included in the plot for clarity).

Appendix B: Post-spallation interferometric measurements

All measurements and analyses reported in Sect. ‘Interferometric Measurements’ were performed for the laser fluence, which was marginally lower than those required for failure initiation. The interferometric data exhibited distinct primary and secondary fringes (see Figs. 8, 9 and 10). Had the spallation initiated because of the primary compressive wave, the optical signal would have contained only the primary fringes, as shown in Fig. 17. The typical wave reflections and mode conversions following spallation are sketched in Fig. 18. The illustrated x-t diagram is the same as shown in Fig. 11, except that the spall plane introduces additional mode-conversions. Primary (p with a numeral subscript) and secondary (s with a numeral subscript) waves are illustrated by the solid and dashed lines, in which the compressive and tensile waves are shown in blue and red, respectively. As shown in the figure, p3 and s2 develop due to the wave reflections from the spall plane. The sketch indicates that the interferometric data recorded at the free surface should contain the signatures of the p1, p3 and s1 waves. By considering the spall depths shown in Table 1 and using 2675 m/s and 2375 m/s as the primary and secondary wave speeds, respectively, the time difference between the arrival of the p3 and s1 waves to the free epoxy surface is calculated to be less than 2 ns. Clearly, it will be difficult to distinguish the signature of the two waves within such a short duration. In addition, the two waves reaching the free surface are of opposite sign (see Fig. 18). Since the primary and secondary waves have comparable magnitudes, their destructive interaction would not allow secondary fringes to develop. Another reason for the diminished interferometric signal is the scattering of the probe beam from the spallation-induced crumbled (and uneven) surface of epoxy.

Fig. 17
figure 17

ad The representative interferometric fringes correspond to the laser fluneces at which the spall had been detected. The data illustrated in (a) to (d) are for 78 μm, 165 μm, 197 μm and 410 μm epoxy film cases, respectively. The associated primary compressive pulses are shown in (a1d1).

Fig. 18
figure 18

The travel and mode conversions of the primary (p) and secondary (s) waves following the development of a spall plane in the epoxy. The red and blue lines indicate tensile and compressive waves, respectively

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Singh, S., Kitey, R. Spall Characterization in Epoxy Via Laser Spallation. Exp Mech 60, 969–985 (2020). https://doi.org/10.1007/s11340-020-00614-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11340-020-00614-4

Keywords

Navigation