Skip to main content
Log in

An enhanced approach for damage detection using the electromechanical impedance with temperature effects compensation

  • Review
  • Published:
Journal of the Brazilian Society of Mechanical Sciences and Engineering Aims and scope Submit manuscript

Abstract

Damage detection is one of the great challenges of the maintenance tasks and it has involved numerous researches to develop techniques in the field of structural health monitoring (SHM). Among different techniques, electromechanical impedance (EMI) technique has attracted attention due to its important and promising results. However, the sensitivity of this technique to variations in environmental conditions can lead to false diagnoses, and the temperature is one of the most critical factors for EMI technique. In view of this point, different researchers have developed compensation techniques to minimize the effects caused by temperature variation in electromechanical impedance measurements. Another important issue related to electromechanical Impedance curves is about the frequency range chosen to be analyzed. Then, the present article introduces an improved approach for damage detection by adding a new step for the temperature compensation technique proposed in a well-established approach in the literature. The proposal comprises a strategy to select the frequency range to compute damage detection indexes, and the technique is demonstrated for an aluminum beam in three different structural conditions: corresponding to the healthy and two types of damaged structure. The results are investigated for four different frequency ranges. The findings demonstrate the effectiveness of the proposed approach to reduce false alarms in damage detection using the EMI technique.

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.

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
Fig. 17
Fig. 18
Fig. 19
Fig. 20
Fig. 21
Fig. 22
Fig. 23

Similar content being viewed by others

References

  1. Abbas S, Li F, Abbas Z, Abbasi TUR, Tu X, Pasha RA (2021) Experimental study of effect of temperature variations on the impedance signature of PZT sensors for fatigue crack detection. Sound Vib 55:1–18

    Google Scholar 

  2. Annamdas VGM, Radhika MA (2013) Electromechanical impedance of piezoelectric transducers for monitoring metallic and non-metallic structures: a review of wired, wireless and energy-harvesting methods. J Intell Mater Syst Struct 24(59):1021–1042. https://doi.org/10.1177/1045389X13481254

    Article  Google Scholar 

  3. Annamdas VGM, Rizzo P (2009) Influence of the excitation frequency in the electromechanical impedance method for SHM applications. In: Meyendorf NG, Peters KJ and Ecke W (eds) Smart sensor phenomena, technology, networks, and systems 2009’, vol 7293. International Society for Optics and Photonics, SPIE, p 72930V. https://doi.org/10.1117/12.815366

  4. Baptista FG, Budoya DE, De Almeida VAD, Ulson JAC (2014) An experimental study on the effect of temperature on piezoelectric sensors for impedance-based structural health monitoring. Sensors 14:1208–1227

    Article  Google Scholar 

  5. Baptista FG, Filho JV (2009) A new impedance measurement system for PZT based structural health monitoring. IEEE Trans Instrum Meas 58(10):3602–3608. https://doi.org/10.1109/TIM.2009.2018693

    Article  Google Scholar 

  6. Baptista FG, Filho JV (2010) Optimal frequency range selection for PZT transducers in impedance-based SHM systems. IEEE Sens J 10(8):1297–1303. https://doi.org/10.1109/JSEN.2010.2044037

    Article  Google Scholar 

  7. Baptista FG, Vieira Filho J, Inman DJ (2010) Influence of excitation signal on impedance-based structural health monitoring. J Intell Mater Syst Struct 21(14):1409–1416. https://doi.org/10.1177/1045389X10385032

    Article  Google Scholar 

  8. Bhalla S (2002) Smart system based automated health monitoring of structures. Master’s thesis, Nanyang Technological University, Singapore

  9. Bhalla S, Soh CK (2004) Electromechanical impedance modeling for adhesively bonded piezo-transducers. J Intell Mater Syst Struct 15(12):955–972. https://doi.org/10.1177/1045389X0404630

    Article  Google Scholar 

  10. Djemana M, Hrairi M, Al Jeroudi Y (2017) Using electromechanical impedance and extreme learning machine to detect and locate damage in structures. J Nondestruct Eval. https://doi.org/10.1007/s10921-017-0417-5

    Article  Google Scholar 

  11. Du F, Wu S, Xu C, Yang Z, Su Z (2021) Electromechanical impedance temperature compensation and bolt loosening monitoring based on modified Unet and multitask learning. IEEE Sens J. https://doi.org/10.1109/JSEN.2021.3132943

    Article  Google Scholar 

  12. Farrar CR, Lieven NAJ, Bement MT (2005) Damage prognosis for aerospace, civil and mechanical systems, 1st edn. Wiley, Hoboken

    Google Scholar 

  13. Giurgiutiu V (2014) Structural health monitoring with piezoelectric wafer active sensors, 2nd edn. Elsevier, Amsterdam

    Google Scholar 

  14. Giurgiutiu V, Rogers CA (1998) Recent advancements in the electromechanical (e/m) impedance method for structural health monitoring and NDE. In: Regelbrugge ME (ed) Smart structures and materials 1998: smart structures and integrated systems, vol 3329. International Society for Optics and Photonics, SPIE, pp 536–547

  15. Grisso BL, Inman DJ (2010) Temperature corrected sensor diagnostics for impedance-based SHM. J Sound Vib 329(12):2323–2336

    Article  Google Scholar 

  16. Huynh T-C, Kim J-T (2018) RBFN-based temperature compensation method for impedance monitoring in prestressed tendon anchorage. Struct Control Health Monit 25(6):1–17. https://doi.org/10.1002/stc.2173

    Article  Google Scholar 

  17. Huynh TC, Dang NL, Kim JT (2018) PCA-based filtering of temperature effect on impedance monitoring in prestressed tendon anchorage. Smart Struct Syst 22(1):57–70

    Google Scholar 

  18. Koo KY, Park S, Lee JJ, Yun CB (2009) Automated impedance-based structural health monitoring incorporating effective frequency shift for compensating temperature effects. J Intell Mater Syst Struct 20(4):367–377. https://doi.org/10.1177/1045389X08088664

    Article  Google Scholar 

  19. Krishnamurthy K, Lalande F, Rogers CA (1996) Effects of temperature on the electrical impedance of piezoelectric sensors. In: Smart structures and materials 1996: smart structures and integrated systems, vol 2717. International Society for Optics and Photonics, SPIE, pp 302–310

  20. Liang C, Sun FP, Rogers CA (1993) Coupled electromechanical analysis of piezoelectric ceramic actuator-driven systems: determination of the actuator power consumption and system energy transfer. Smart Struct Mater. https://doi.org/10.1117/12.152767

    Article  Google Scholar 

  21. Liang C, Sun FP, Rogers CA (1996) Electro-mechanical impedance modeling of active material systems. Smart Mater Struct 5(2):171–186. https://doi.org/10.1088/0964-1726/5/2/006

    Article  Google Scholar 

  22. Lim HJ, Kim MK, Sohn H, Park CY (2011) Impedance based damage detection under varying temperature and loading conditions. Ndt E Int 44:740–750

    Article  Google Scholar 

  23. Min J, Park S, Yun C-B (2010) Impedance-based structural health monitoring using neural networks for autonomous frequency range selection. Smart Mater Struct 19(12):125011. https://doi.org/10.1088/0964-1726/19/12/125011

    Article  Google Scholar 

  24. Na WS, Baek J (2018) A review of the piezoelectric electromechanical impedance based structural health monitoring technique for engineering structures. Sensors 18(5):1307

    Article  Google Scholar 

  25. Park G, Kabeya K, Cudney HH, Inman DJ (1998) Removing effects of temperature changes from piezoelectric impedance-based qualitative health monitoring. In: Smart structures and materials 1998: sensory phenomena and measurement instrumentation for smart structures and materials, vol 3330, SPIE. https://doi.org/10.1117/12.316963

  26. Park G, Kabeya K, Cudney HH, Inman DJ (1999) Impedance-based structural health monitoring for temperature varying applications. JSME Int J Ser A 42(2):249–258. https://doi.org/10.1299/jsmea.42.249

    Article  Google Scholar 

  27. Park G, Sohn H, Farrar CR, Inman DJ (2003) Overview of piezoelectric impedance-based health monitoring and path forward. Shock Vib Dig 35(6):451–463. https://doi.org/10.1177/05831024030356001

    Article  Google Scholar 

  28. Peairs DM, Tarazaga PA, Inman DJ (2007) Frequency range selection for impedance-based structural health monitoring. J Vib Acoust 129(6):701–709. https://doi.org/10.1115/1.2775506

    Article  Google Scholar 

  29. Perera R, Torres L, Díaz FJ, Barris C, Baena M (2021) Analysis of the impact of sustained load and temperature on the performance of the electromechanical impedance technique through multilevel machine learning and FBG sensors. Sensors 21(17):5755

    Article  Google Scholar 

  30. Selva P, Cherrier O, Budinger V, Lachaud F, Morlier J (2013) Smart monitoring of aeronautical composites plates based on electromechanical impedance measurements and artificial neural networks. Eng Struct 56:794–804

    Article  Google Scholar 

  31. Sepehry N, Shamshirsaz M, Abdollahi F (2011) Temperature variation effect compensation in impedance-based structural health monitoring using neural networks. J Intell Mater Syst Struct 22:1975–1982. https://doi.org/10.1177/1045389X11421814

    Article  Google Scholar 

  32. Singh SK, Malinowski PH (2022) An innovative data-driven probabilistic approach for damage detection in electromechanical impedance technique. Compos Struct 295:115808

    Article  Google Scholar 

  33. Sun FP, Chaudhry ZA, Rogers CA, Majmundar M, Liang C (1995b) Automated real-time structure health monitoring via signature pattern recognition. In: Smart structures and materials 1995: smart structures and integrated systems, vol 2443. SPIE, pp 236–247

  34. Sun FP, Chaudhry Z, Liang C, Rogers CA (1995) Truss structure integrity identification using PZT sensor-actuator. J Intell Mater Syst Struct 6(1):134–139. https://doi.org/10.1177/1045389X9500600117

    Article  Google Scholar 

  35. Tseng KK, Naidu A (2002) Non-parametric damage detection and characterization using smart piezoceramic material. Smart Mater Struct 11(3):317–329. https://doi.org/10.1088/0964-1726/11/3/301

    Article  Google Scholar 

  36. Wang D, Song H, Zhu H (2015) Electromechanical impedance analysis on piezoelectric smart beam with a crack based on spectral element method. Math Probl Eng. https://doi.org/10.1155/2015/713501

    Article  Google Scholar 

  37. Worden K, Dulieu-Barton JM (2004) An overview of intelligent fault detection in systems and structures. Struct Health Monit 3(1):85–98. https://doi.org/10.1177/1475921704041866

    Article  Google Scholar 

  38. Xu G, Xu B, Xu C, Luo Y (2016) Temperature effects in the analysis of electromechanical impedance by using spectral element method. Multidiscip Model Mater Struct 12(1):119–132. https://doi.org/10.1108/MMMS-03-2015-0015

    Article  Google Scholar 

  39. Zhou SW, Liang C, Rogers CA (1996) An impedance-based system modeling approach for induced strain actuator-driven structures. J Vib Acoust 118:323–331. https://doi.org/10.1115/1.2888185

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Lorena Lopes Dias.

Ethics declarations

Conflict of interest

The authors have no relevant financial or non-financial interests to disclose.

Additional information

Technical Editor: Adriano Fagali de Souza.

Publisher's Note

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

The original online version of this article has been corrected since the figures and Tables were misplaced during corrections.

Appendices

Appendix 1 Overview on the Park et al. [26] compensation technique

This appendix shows the EMI curves when applying the Park et al. [26] temperature compensation technique. Figure 24 shows the real part of the EMI obtained for the aluminum beam in the undamaged condition in two different temperatures (10 and 30 \(^{\circ }\)C) and 4 different stages of the method (i.e., iterations). The values of \(V_a\) and \(\delta ^S\) for each iteration are shown in Table 4. After the first iteration the difference in the vertical axis reduced by summing \(\delta ^S\) to the curve. This parameter presents a small variation even after multiple iterations, as shown in Table 4. This characteristics is verified because \(\delta ^S\) is computed by the difference between \(Z_R\) and the measured signal shifted horizontally.

Park et al. [26] verified that it is possible to detect damages in the incipient phases, even considering temperature variations from 25 to 75 \(^{\circ }\)C, with a step of 12.5 \(^{\circ }\)C. The experiments were carried out in different frequency ranges according to the monitored structure: carbon steel beam (70–80 kHz), bolted pipe joint (70–80 kHz), gears (190–220 kHz), and composite-reinforced structure (54–63 kHz).

Table 4 Values of \(V_a\) and \({\delta ^{S}}\) in each iteration
Fig. 24
figure 24

Real part of electromechanical impedance at different iterations of Park et al. [26] technique: 10 °C (black continuous line) and 30 °C (blue-dashed line)

Appendix 2 Results for the frequency ranges numbers 1 and 2 (Table 3)

Figures 25, 26 and 27 show the EMI signal after applying the approach proposed by Park et al. [26]. These figures contain, respectively, the signals for the undamaged condition, with damage 1 (added mass) and damage 2 (mechanical cut) for two different frequency ranges, corresponding to the bandwidths 15 and 30 kHz. The values of the difference between the RMSD and CCDM of the signals for both damaged and undamaged conditions are shown in Figs. 28 and 29. Figure 30 shows similar results regarding damage detection (i.e., no false negatives) for the frequency range number 2 (bandwidth 30 kHz) after applying the new approach. On the other hand, Appendix 3 shows that this approach can not be successfully applied by considering the frequency range number 1 because a stable value of S is not achieved.

Fig. 25
figure 25

Real part of the electromechanical impedance (undamaged condition) using Park et al. [26] compensation technique for two frequency bandwidths: 15 and 30 kHz. Curves for the temperatures \(-10\,^{\circ }\)C, 4 \(^{\circ }\)C, 24 \(^{\circ }\)C (reference), 36 \(^{\circ }\)C, 52 \(^{\circ }\)C, 64 \(^{\circ }\)C and 80 \(^{\circ }\)C

Fig. 26
figure 26

Real part of the electromechanical impedance (damage 1) using Park et al. [26] compensation technique for two frequency bandwidths: 15 and 30 kHz. Curves for the temperatures \(-10\,^{\circ }\)C, 5 \(^{\circ }\)C, 25 \(^{\circ }\)C, 35 \(^{\circ }\)C, 50 \(^{\circ }\)C, 65 \(^{\circ }\)C and 80 \(^{\circ }\)C

Fig. 27
figure 27

Real part of the electromechanical impedance (damage 2) using Park et al. [26] compensation technique for two frequency bandwidths: 15 and 30 kHz. Curves for the temperatures \(-10\,^{\circ }\)C, 5 \(^{\circ }\)C, 25 \(^{\circ }\)C, 35 \(^{\circ }\)C, 50 \(^{\circ }\)C, 65 \(^{\circ }\)C and 80 \(^{\circ }\)C

Fig. 28
figure 28

Difference between the RMSD values computed for the signals in the damaged and undamaged conditions compensated by Park et al. [26] technique for the FR numbers 1 and 2. Yellow slanted bar hatch depicts damage 1 and blue bar with horizontal hatch represents damage 2

Fig. 29
figure 29

Difference between the CCDM values computed for the signals in the damaged and undamaged conditions compensated by Park et al. [26] technique for the FR numbers 1 and 2. Yellow slanted bar hatch depicts damage 1 and blue bar with horizontal hatch represents damage 2

Fig. 30
figure 30

Difference between the RMSD and CCDM values computed for the signals in the damaged and undamaged conditions after applying the new approach for the FR number 2. Yellow slanted bar hatch depicts damage 1 and blue bar with horizontal hatch represents damage 2

Appendix 3 Analysis of the S-curve

Table 5 shows the values of S for each frequency range. Note that it is numerically approximately a constant value (around \(10^3\)) for the frequency ranges 2, 3 and 4. However, its variation is more significant for the frequency range number 1.

This behavior suggests that the approach can not be successfully applied when using this frequency range. Then, the evaluation of the parameter S is an important part of this proposed strategy to select the frequency range.

Table 5 S and \(N_r\) values for the four frequency ranges (FR)

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Dias, L.L., Lopes, K.W., Bueno, D.D. et al. An enhanced approach for damage detection using the electromechanical impedance with temperature effects compensation. J Braz. Soc. Mech. Sci. Eng. 45, 228 (2023). https://doi.org/10.1007/s40430-023-04127-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s40430-023-04127-8

Keywords

Navigation