Elsevier

Materials & Design

Volume 53, January 2014, Pages 236-249
Materials & Design

Effect of corrosion on the fatigue life and fracture mechanisms of 6101 aluminum alloy wires for car manufacturing applications

https://doi.org/10.1016/j.matdes.2013.06.079Get rights and content

Highlights

  • We examine the fatigue–corrosion behavior of an Al–Mg–Si 6101 alloy.

  • The fatigue endurance of this alloy was studied using the stress-life approach.

  • Three metallurgical states representative of three cold-drawing steps were considered.

  • Fatigue–corrosion interactions modify the failure mode.

  • Hydrogen embrittlement mechanism could occur during fatigue–corrosion tests.

Abstract

An innovative solution for the automotive industry is to replace the copper used for wiring harnesses with aluminum alloys, such as the aluminum–magnesium–silicon 6101 alloy. Wiring harnesses are composed of thin strand arms obtained by a wire drawing process. These strands are susceptible to exposure to a corrosive environment and fatigue solicitations simultaneously. The fatigue endurance of this alloy was studied using the stress-life approach for three metallurgical states representative of three cold-drawing steps. Fatigue tests performed in corrosive media tests highlighted a strong decrease of the 6101 alloy lifetime due to fatigue–corrosion interactions and a modification of failure modes.

Introduction

Considering the rarefaction of fossil fuels and the need to reduce greenhouse gas emissions, automotive manufacturers are looking for innovative solutions to reduce both the cost and weight of wiring harnesses. One promising solution is the substitution of copper with aluminum alloys, such as the aluminum–magnesium–silicon 6101 alloy, which is currently used for high-power electrical applications. Wiring harnesses require thin strand arms, which are obtained by wire-drawing. This metalworking process consists of a reduction in the rod or wire diameter by passing it through a series of drawing dies interspersed, if necessary, with aging heat treatments to release the stresses induced by the process. The different microstructures obtained after each step of the manufacturing process strongly depend on the thermo-mechanical process characteristics, which also affect the mechanical properties and the corrosion behavior of the material. The strands are susceptible to exposure to several corrosive environments, e.g., de-icing road salt, coolant, and windshield washer fluids, and fatigue solicitations due to motor and engine vibrations. These combined damages can lead to a premature degradation of wiring harnesses. Many studies have focused on the role of pre-corrosion damage on the residual fatigue life and the effect of localized corrosion defects on fatigue crack initiation in aluminum alloys. Most such studies have focused on the 2xxx and 7xxx series and fatigue crack initiation from pitting corrosion defects [1], [2], [3], [4], [5], [6], [7], [8], [9], [10], [11]. These alloys have shown a large reduction in fatigue life due to corrosion pits, providing potential stress concentrator sites at which fatigue cracks can initiate. The pit size (average depth and width) seems to be an essential parameter; the fatigue life is successfully predicted by 2D AFGROW (Fracture Mechanics and Fatigue Crack Growth Analysis software tool) calculations based on initial crack sizes similar to those of pits [4]. Dolley et al. have shown, for AA 2024-T3, that the fatigue lives were reduced by more than one order of magnitude after 384 h of pre-corrosion in a 0.5 M NaCl solution compared to those of uncorroded specimens. The reduction of fatigue life was related to the time of exposure to the corrosive environment and pit size [11]. Burns et al. have shown that the fatigue life of a 7075-T6511 alloy was reduced substantially by EXCO pre-corrosion, but this was nearly independent of exposure time after initial-sharp degradation, scaling with the evolution of pit-cluster size and initial stress intensity range [12]. Liao et al. have studied the remaining fatigue life of 7075-T6511 aircraft wing skins containing natural exfoliation corrosion. It was found that the relative humidity did not have a significant effect on the fatigue life of naturally exfoliated specimens. Exfoliation above a “critical” level (undetermined in the paper of Liao et al. due to the limited number of tests) could significantly decrease the fatigue life of the naturally exfoliated specimens. When the exfoliation damage is not very severe, the discontinuities present in the fillet of the samples, where the stress concentration Kt was equal to 1.23, could override the exfoliation damage and become the primary crack origin. Above the “critical” level, the cracks primarily nucleated from corrosion pits and intergranular corrosion defects [13]. Pauze has studied fatigue crack initiation on localized corrosion defects for a 2024-T351 aluminum alloy, i.e., intergranular corrosion defects and pits. Both defects acted as preferential initiation sites for transgranular fatigue cracks [14].

Finally, the fatigue life duration for pre-corroded samples is essentially related to the fatigue crack initiation step on preexisting localized corrosion defect. A possible effect of H-uptake during pre-corrosion treatment on crack propagation rates is possible but it is probably negligible compared to the strong decrease of the duration of the crack initiation step.

However, for the tests performed directly in the corrosive media, the same conclusion cannot be advanced seeing that corrosion and fatigue mechanisms could interact as well during crack initiation as during crack propagation.

Most of literature results about fatigue–corrosion interactions have shown that the reduction in the fatigue life in NaCl solutions could be attributed to premature crack initiation from surface corrosion defects [15], [16], [17].

The characteristics of these localized corrosion defects such as pit geometry [18], [19], micro-topographic strain concentration and local stress distribution around pit [20] are essential in determining the pit-to-crack transition such as fatigue life tests performed on pre-corroded samples. Mutombo and du Toit have characterized the fatigue–corrosion endurance behavior of a welded 6061-T651 aluminum alloy in a 3.5% (by weight) NaCl solution at various applied stress amplitudes [21]. The reduction in the fatigue life in the corrosive solution compared to the fatigue life in air was also most strongly related to the presence of pits, which nucleated on second-phase particles or precipitates and acted as preferential fatigue crack initiation sites. Pauze [14] has shown for a 2024 aluminum alloy that fatigue–corrosion interactions only acted on the crack propagation regime. The crack initiation occurred on corrosion defects which were strictly the same as those obtained during continuous immersion in corrosive media without applied stress. The effect of fatigue corrosion synergy only occurred during the crack propagation with the competition between the growth rate of the transgranular fatigue–corrosion crack and the growth rate of the intergranular stress corrosion crack. The propagation rate of the intergranular corrosion defect whose kinetic is relatively important during the first hours is quickly followed by a slow transgranular fatigue–corrosion propagation. The effect of the corrosive media is relatively low (factor 2–5) during this second step and occurs for low ΔK values. Finally transgranular fatigue–corrosion propagation is comparable to the propagation obtained in air [14].

In the present work, the reference metal was a 9.5 mm wire rod of AA 6101 (T4 metallurgical state); this rod was cold-drawn to manufacture the strand arms. The steps of the cold-drawing process used in the industry are summarized in Fig. 1. An intermediate aging heat treatment (185 °C–10 h) following the cold-drawing from a 9.5 mm diameter wire rod to a 1.34 mm diameter strand was applied. Finally, a final cold-drawing step provides a thin strand with a 0.51 mm final diameter (T9 metallurgical state). The influence of a pre-corrosion step on the fatigue and fatigue–corrosion behaviors of both T4 (the initial material) and T9 (the metallurgical state used) samples were studied. In parallel, some T4 samples were aged at 185 °C for 10 h in the laboratory to specifically study the influence of the intermediate heat treatment (T4 + 185 °C–10 h) on the material properties.

In a previous work [22], the corrosion behavior of the aluminum–magnesium–silicon 6101 alloy has been studied in chloride solution in terms of the different metallurgical states resulting from the wire drawing process (T4, T4 + 185 °C–10 h, and T8 metallurgical states; T8 corresponds to T4 + cold drawing until a 1.34 mm-diameter rod was obtained, followed by the 185 °C–10 h aging treatment; T9 corresponds to the same procedure as T8 except with an additional cold drawing step, which does not modify the corrosion morphology observed compared to the T8 state). Regardless of its metallurgical state, AA 6101 was found to be susceptible to localized corrosion. Scanning electron microscopy (SEM) observations of the electrode surfaces obtained after potentiokinetic polarization tests have shown that all metallurgical states were susceptible to two types of localized corrosion morphologies, i.e., pitting corrosion (Fig. 2(a) and (d)) and matrix dissolution around particles (Fig. 2(b)). Many thin corrosion filaments were observed around and between pit cavities according to literature results (Fig. 2(a) and (d)) [23]. These observations are consistent with the literature results. The main intermetallics present in Al–Mg–Si–Fe alloys are Fe-rich particles such as Al3Fe, α-Al8Fe2Si, β-Al5FeSi and π-Al8FeMgSi6 [24], [25], [26] and Mg2Si precipitates [27].

Pits in Al–Mg–Si alloys often initiate on intermetallic particles. Blanc et al. [28] have demonstrated using interferometry and surface observations that Mg2Si particles suffer preferential Mg dissolution just after a few seconds of immersion in aggressive media, in agreement with other authors [29], [30]. Preferential initiation for deep corrosion was then observed at Mg2Si sites [28]. Concerning Al–Fe–Si precipitates, they are known to be cathodic respective to the aluminum matrix [26], [31]. During immersion in an aggressive media, the galvanic cells created between these intermetallics and the matrix initiate the dissolution of the surrounding matrix [32].

For the T4 + 185 °C–10 h metallurgical state only, observations of the electrode surfaces revealed a third form of localized corrosion, i.e., intergranular corrosion (Fig. 2(c)). As a reminder, for T4 + 185 °C–10 h, numerous precipitates arrange into a continuous film, with needle (average dimension = 35 nm × 140 nm) or globular shapes observed at the grain boundaries, unlike in other metallurgical states, which could explain the susceptibility to intergranular corrosion of the T4 + 185 °C–10 h samples. Therefore, the results showed that the morphology of the corrosion defects depended on the metallurgical state of the 6101 aluminum alloy, suggesting different fatigue behavior after pre-corrosion damage or during fatigue–corrosion tests.

In this paper, the effect of pre-corrosion damage and fatigue–corrosion synergistic impact on fatigue life was studied for T4, T4 + 185 °C–10 h, and T9 samples. Considerable effort was applied to the SEM analyses of the fracture surfaces resulting from the different tests for each metallurgical state; the observations were explained based on the microstructure and corrosion susceptibility of the different metallurgical states.

Section snippets

Material

In the present work, AA 6101 in the T4 metallurgical state was provided by Rio Tinto Alcan as a wire rod with a 9.5 mm diameter. Its chemical composition is given in Table 1. The T9 samples were prepared by SASA Trefil’Alu, whereas the T4 + 185 °C–10 h samples were prepared in the laboratory.

The microstructural characteristics of the initial wire rod and of the different metallurgical states obtained by the wire drawing process have been characterized in a previous work [22] by optical microscopy

Tensile behavior

The tensile curves obtained for the three metallurgical states are presented in Fig. 5. The T9 metallurgical state presented a strong increase in the yield strength and ultimate tensile strength combined with a very low plastic deformation. This behavior corresponded well with a Hall-Petch strengthening effect related to the very fine grain size due to the wire-drawing process and the high dislocation density observed by TEM.

Concerning the T4 and T4 + 185 °C–10 h metallurgical states, the heat

Conclusions

The fatigue behavior of AA 6101 under T4, T4 + 185 °C–10 h, and T9 metallurgical states was studied in air for healthy and pre-corroded samples. The results were compared to those obtained during fatigue–corrosion tests in NaCl solutions.

In air, the heat treatment led to a strong decrease in the fatigue endurance of the material due to the formation of PFZ and Mg2Si intergranular precipitates, which promoted intergranular failure. The complete wire drawing process seemed to increase the fatigue

Acknowledgments

This work was financially supported by LEONI Wiring Systems, France. The material was provided by Rio Tinto Alcan and cold drawn by SASA Trefil’Alu. The authors thank Didier Albert (Rio Tinto Alcan), Louis Chretien (LEONI Wiring Systems), and Jean-Jacques Zyla (SASA Trefil’Alu). The authors also thank Marie-Christine Lafont for the TEM observations.

References (64)

  • G.S. Chen et al.

    Transition from pitting to fatigue crack growth-modeling of corrosion fatigue crack nucleation in a 2024-T3 aluminum alloy

    Mater Sci Eng A

    (1996)
  • U. Zupanc et al.

    Effect of pitting corrosion on fatigue performance of shot-peened aluminium alloy 7075-T651

    J Mater Process Technol

    (2010)
  • Q.Y. Wang et al.

    Effect of pitting corrosion on very high cycle fatigue behaviour

    Scripta Mater

    (2003)
  • D.L. DuQuesnay et al.

    Fatigue crack growth from corrosion damage in 7075-T6511 aluminium alloy under aircraft loading

    Int J Fatigue

    (2003)
  • S. Kim et al.

    Fatigue crack formation and growth from localized corrosion in Al–Zn–Mg–Cu

    Eng Fract Mech

    (2009)
  • S.I. Roklin et al.

    Effect of pitting corrosion on the fatigue crack initiation and fatigue life

    Eng Fract Mech

    (1999)
  • K. Van der Walde et al.

    Initiation and shape development of corrosion-nucleated fatigue cracking

    Int J Fatigue

    (2007)
  • A.T. Kermanidis et al.

    Fatigue and damage tolerance behaviour of corroded 2024 T351 aircraft aluminum alloy

    Theory Appl Fract Mech

    (2005)
  • K.M. Gruenberg et al.

    Predicting fatigue life of pre-corroded 2024–T3 aluminum from breaking load tests

    Int J Fatigue

    (2004)
  • P. Shi et al.

    Corrosion fatigue and multiple site damage reliability analysis

    Int J Fatigue

    (2003)
  • E.J. Dolley et al.

    The effect of pitting corrosion on fatigue life

    Fatigue Fract Eng Mater

    (2000)
  • J.T. Burns et al.

    Effect of corrosion severity on fatigue evolution in Al–Zn–Mg–Cu

    Corros Sci

    (2010)
  • M. Liao et al.

    Fatigue modeling for aircraft structures containing natural exfoliation corrosion

    Int J Fatigue

    (2007)
  • Pauze N. Ph.D. thesis: Fatigue corrosion dans le sens travers court de tôles d’aluminium 2024-T351 présentant des...
  • R.M. Chlistovsky et al.

    Corrosion-fatigue behaviour of 7075-T651 aluminum alloy subjected to periodic overloads

    Int J Fatigue

    (2007)
  • M. Rebiere et al.

    Corrosion fatigue mechanisms of an 8090 Al–Li–Cu alloy

    Mater Sci Eng A

    (1990)
  • S.I. Rokhlin et al.

    Effect of pitting corrosion on fatigue crack initiation and fatigue life

    Eng Fract Mech

    (1999)
  • D.A. Horner et al.

    Novel images of the evolution of stress corrosion cracks from corrosion pits

    Corros Sci

    (2011)
  • P.S. Pao et al.

    On fatigue crack initiation from corrosion pits in 7075-T7351 aluminum alloy

    Scripta Mater

    (2000)
  • J.T. Burns et al.

    Driving forces for localized corrosion-to-fatigue crack transition in Al–Zn–Mg–Cu

    Fatigue Fract Eng Mater

    (2011)
  • K. Mutombo et al.

    Corrosion fatigue behaviour of aluminium alloy 6061-T651 welded using fully automatic gas metal arc welding and ER5183 filler alloy

    Int J Fatigue

    (2011)
  • A. Laurino et al.

    Effect of the thermomechanical history on the corrosion behavior of 6101 aluminum–magnesium–silicon–iron alloy in NaCl solutions

    ECS Trans

    (2012)
  • R. Ambat et al.

    Effect of iron-containing intermetallic particles on the corrosion behaviour of aluminium

    Corros Sci

    (2006)
  • L.F. Mondolfo

    Aluminum alloys: structure and properties

    (1976)
  • Y.L. Liu et al.

    The complex microstructures in an as-cast Al–Mg–Si alloy

    Mater Lett

    (1999)
  • S. Karabay et al.

    Investigation of extrusion ratio effect on mechanical behavior of extruded alloy AA-6101 from the billets homogenised-rapid quenched and as-cast conditions

    J Mater Process Technol

    (2005)
  • R.G. Buchheit

    A compilation of corrosion potentials reported for intermetallic phases in aluminum alloys

    J Electrochem Soc

    (1995)
  • C. Blanc et al.

    Application of phase shifting interferometric microscopy to studies of the behavior of coarse intermetallic particles in 6056 aluminum alloy

    Corros Sci

    (1998)
  • F. Eckermann et al.

    The influence of MgSi particle reactivity and dissolution processes on corrosion in Al–Mg–Si alloys

    Electrochim Acta

    (2008)
  • K. Mizuno et al.

    Surface reactions during pickling of an aluminum–magnesium–silicon alloy in phosphoric acid

    Corros Sci

    (2001)
  • Z. Szklarska-Smialowska

    Pitting corrosion of aluminum

    Corros Sci

    (1999)
  • H. Ezuber et al.

    A study on the corrosion behavior of aluminum alloys in seawater

    Mater Des

    (2008)
  • Cited by (0)

    View full text