Research paper
Experimental and numerical analysis of the heat generated by plastic deformation in quasi-static uniaxial tensile tests

https://doi.org/10.1016/j.mechmat.2020.103398Get rights and content

Highlights

  • Experimental and numerical analysis of the temperature rise in the quasi-static uniaxial tensile tests.

  • The predicted temperature rise is almost independent of the selected heat transfer coefficients.

  • Comparison between the Zehnder model and the Aravas model.

  • Taylor–Quinney coefficient increases as the plastic deformation accumulates for the aluminium alloy AA6016-T4.

Abstract

This study evaluates the temperature variation observed in quasi-static uniaxial tensile tests, due to the heat generated by plastic deformation. The AA6016-T4 aluminium alloy was the material selected, considering different values of crosshead velocity (from 0.01 mm/s up to 1 mm/s). The temperature variation was also evaluated during the stress relaxation test. A finite element model of the uniaxial tensile test is presented, which takes into account the heat generated by plastic deformation, as well as the effect of the heat losses to the environment (convective heat transfer coefficient) and to the grips (interfacial heat transfer coefficient). The numerical results show that the predicted temperature variation is almost independent of the selected heat transfer coefficients. On the other hand, the temperature rise is influenced by the Taylor–Quinney coefficient. The comparison between experimental and numerical temperatures shows that the Zehnder model (increasing Taylor–Quinney coefficient) provides more accurate results than the Aravas model (decreasing Taylor–Quinney coefficient). Nevertheless, the evolution of the Taylor–Quinney coefficient defined by the Zehnder model assumes a constant value for the hardening coefficient, which does not fit the hardening behaviour observed for this aluminium alloy.

Introduction

In sheet metal forming processes, the blank temperature increases during the forming operation due to the heat generated by plastic deformation and by friction between the blank and the forming tools (Pereira and Rolfe, 2014). This is particularly relevant in mass production lines due to the periodic input of heat into the forming tools. In fact, the temperature variation produces changes both in the mechanical behaviour of the blank and particularly in the lubrication conditions, contributing to process variability (Hazra et al., 2011). Therefore, this phenomenon should be considered during the optimization of the process parameters, mainly in case of high strength steels and for high speed forming operations. Hence, the accurate numerical modelling of these forming processes requires a thermo-mechanical analysis, to account for the influence of the heat generated by plastic deformation.

Since the deformation process of metallic materials in the elastic regime leads to insignificant volume changes (lower than 1%), the temperature variation defined by the Joule–Thomson effect for solids is always slight (Kuo et al., 2005). In contrast, the plastic deformation generates a significant amount of heat (Farren and Taylor, 1925), which can lead to substantial temperature increases if the deformation process is fast (no time for heat exchange). Indeed, during plastic deformation, the mechanical energy is predominantly dissipated as heat, while the stored or latent energy remains in the material after removing the external loads (Hodowany et al., 2000). The first attempts to determine experimentally the fraction of plastic work converted into heat was carried by Taylor and Quinney (1934), using calorimetric methods. They found that the fraction of dissipated energy (also known as Taylor–Quinney coefficient) is about 90% for different metallic materials.

Some experimental studies show that the Taylor–Quinney coefficient may depend on the plastic strain and on the plastic strain rate. The study performed by Macdougall (2000) on an aluminium alloy shows that the Taylor–Quinney coefficient increases with the plastic strain, ranging roughly from 0.5 to 0.9, which is in agreement with the model developed by Zehnder (1991). An identical tendency was found by Rusinek and Klepaczko (2009) for TRIP steels, taking into account the phase transformation of austenite into martensite during plastic deformation. On the other hand, the influence of the plastic strain rate on the Taylor–Quinney coefficient was experimentally studied by Zhang et al. (2017), using the Kolsky bar with an infrared temperature measurement system. For the 7075-T651 aluminium alloy, they found that the Taylor–Quinney coefficient increases from 0.4 to 0.9 as the strain rate increases from 1100 to 4200 s−1. Nonetheless, a different conclusion was obtained by Hodowany et al. (2000) for the 2024-T3 aluminium alloy, showing that the fraction of plastic work dissipated as heat is not sensitive to the strain rate over a range from 1 s−1 to 3000 s−1. Recently, the dependence of the Taylor–Quinney coefficient on the dynamic loading mode (tension, compression and shear) was experimentally studied by Rittel et al. (2017). The results show that identical coefficients are measured in tension, compression and shear, except for the commercially pure titanium, which is known to exhibit tension−compression asymmetry in its mechanical properties (Revil-Baudard et al., 2015).

In order to take advantage of the adiabatic conditions, high strain rate techniques (e.g. the Kolsky pressure bar) are commonly adopted to calculate the fraction of plastic work converted into heat, where the amount of heat generated is captured by infrared temperature measurements (Mason et al., 1994). Alternatively, hybrid approaches have been developed, combining experimental measurements with numerical analysis through inverse analysis (Zehnder et al., 1998). These approaches are particularly attractive in mechanical tests involving low and intermediate strain rates, i.e. loading under non-adiabatic conditions. The hybrid method developed by Pottier et al. (2013) combines full field measurements (strains and temperature) on heterogeneous tensile tests with numerical results of the thermo-mechanical finite element analysis. They use a Levenberg–Marquardt algorithm to optimize sequentially the mechanical parameters and the Taylor–Quinney coefficient.

The experimental techniques used for temperature measurement can be categorized into two groups: (i) contact and (ii) non-contact techniques. The temperature variations induced by plastic deformation are commonly measured using thermocouples (contact technique) or infrared thermography (non-contact technique) (Ait-Amokhtar et al., 2008). Compared to thermocouples, infrared thermography has the advantage of providing the thermal image of the surface under investigation, while thermocouples measure the temperature on a single point. Therefore, infrared thermography is commonly used to study local temperature variations, such as the ones associated with the Portevin–Le Chatelier effect of Al–Mg alloys (Ait-Amokhtar et al., 2008; Bernard et al., 2013). On the other hand, thermocouples can be inserted in the specimen, allowing the temperature measurement in hidden regions, which is very useful for some experimental studies (Pereira and Rolfe, 2014). Thermocouples spot welded on the specimen surface were reported to present a response time of about 10 milliseconds (Pandey and Chand, 2003), but response times of 10 microseconds can be attained in case of thermocouples embedded in the specimen (Rittel, 1998). Infrared thermography cameras have a response time between milliseconds and microseconds, for cameras based on thermal detectors and on quantum detectors, respectively (Carlomagno and Cardone, 2010). Finally, although the temperature measurement with both systems depends on the working conditions, infrared thermography also requires the calibration of the surface emissivity, which can change during plastic deformation (Hodowany et al., 2000). In order to minimize emissivity problems, the specimen surface is often painted with a high-emissivity graphic-black coating (Ait-Amokhtar et al., 2008; Bernard et al., 2013).

The purpose of this paper is to use temperature variation measurements from quasi-static uniaxial tensile tests to quantify the fraction of plastic work converted into heat. The accuracy of the numerical models that take into account the Taylor–Quinney coefficient as a function of plastic strain is evaluated. The material and the experimental setup of the quasi-static uniaxial tensile tests used are described in Section 2, while the experimental results are presented in Section 3. The proposed thermo-mechanical finite element model is presented in Section 4 including, both the heat generation and the heat losses that occur during the test. The comparison between numerical and experimental temperature results is performed in Section 5, highlighting the influence of the Taylor–Quinney coefficient on the predicted temperature. Section 6 contains the main conclusions of this study.

Section snippets

Material

The age-hardenable aluminium alloy AA6016-T4 was adopted for this study, in sheet format with a thickness of 1.05 mm. The chemical composition of the alloy AA6016-T4 is presented in Table 1. The designation T4 is associated with the following sequence: solution heat treatment (SHT), rapid quenching and natural aging to a substantially stable condition. The mechanical behaviour of the alloy was evaluated after 1, 7 and 18 months of natural aging by Simões et al. (2019). The results show that the

Uniaxial tensile tests

The stress–strain curves obtained from the uniaxial tensile tests performed at crosshead velocities of 0.01, 0.1 and 1 mm/s, with the specimens oriented along the RD and aging times of 1 month and 18 months, are presented in Fig. 4. The increase of the natural aging time leads to an increase in the alloy strength (Simões et al., 2019). Moreover, the increase of the crosshead velocity leads to negligible differences in the flow stress curves. Therefore, this aluminium alloy can be assumed strain

Finite element model

Despite the uniaxial tensile tests are carried out at room temperature, the specimen temperature increases due to the heat generated by plastic deformation. Therefore, numerical simulation should address the thermo-mechanical analysis of the process (Martins et al., 2017). The numerical simulations were carried out with the in-house finite element code DD3IMP (Menezes and Teodosiu, 2000), which has been developed to simulate sheet metal forming processes (Oliveira et al., 2008). Only an eighth

Uniaxial tensile tests

The temperature in the specimen volume is defined by the balance between the amount of heat generated, the heat conduction and the heat losses. The heat losses are first analysed in this section, namely the ones due to natural convection and due to contact conductance with the grips, which are defined by the natural convective heat transfer coefficient (hair) and the interfacial heat transfer coefficient (hgrip), respectively. Regarding the natural convective heat transfer coefficient, the

Conclusions

This study presents the analysis of the heat generated by plastic deformation in quasi-static uniaxial tensile tests, comparing experimental and numerical results. The age-hardenable aluminium alloy AA6016-T4 is adopted in the present study, comparing two different values of natural aging time, namely 1 month and 18 months. Although the adopted alloy is strain rate insensitive at room temperature, three different values of crosshead velocity (0.01 mm/s, 0.1 mm/s and 1 mm/s) are used. This

CRediT authorship contribution statement

D.M. Neto: Conceptualization, Methodology, Investigation, Writing - original draft. V.M. Simões: Validation, Investigation, Formal analysis, Writing - original draft. M.C. Oliveira: Formal analysis, Writing - review & editing, Investigation. J.L. Alves: Methodology, Software. H. Laurent: Resources, Writing - review & editing, Supervision. A. Oudriss: Investigation, Validation. L.F. Menezes: Software, Supervision.

Declaration of Competing Interest

Manuscript entitled “Experimental and numerical analysis of the heat generated by plastic deformation in quasi-static uniaxial tensile tests”, by D.M. Neto, V.M. Simões, M.C. Oliveira, J.L. Alves, H. Laurent, A. Oudriss and L.F. Menezes. The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

Acknowledgements

The authors would like to acknowledge the funding from of the Foundation for Science and Technology (FCT) under projects PTDC/EMS-TEC/6400/2014 (POCI-01-0145-FEDER-016876), PTDC/EMS-TEC/0702/2014 (POCI-01-0145-FEDER-016779) and PTDC/EME-APL/29713/2017 (CENTRO-01-0145-FEDER-029713) by UE/FEDER through the program COMPETE 2020. The support under the project MATIS (CENTRO-01-0145-FEDER-000014) and UID/EMS/00285/2020 is also acknowledged.

References (51)

  • J.J. Mason et al.

    On the strain and strain rate dependence of the fraction of plastic work converted to heat: an experimental study using high speed infrared detectors and the Kolsky bar

    Mech. Mater.

    (1994)
  • L.F. Menezes et al.

    Three-dimensional numerical simulation of the deep-drawing process using solid finite elements

    J. Mater. Process. Technol.

    (2000)
  • D.M. Neto et al.

    Thermo-mechanical finite element analysis of the AA5086 alloy under warm forming conditions

    Int. J. Solids Struct.

    (2018)
  • K.N. Pandey et al.

    Deformation based temperature rise: a review

    Int. J. Press Vessel Pip

    (2003)
  • M.P. Pereira et al.

    Temperature conditions during “cold” sheet metal stamping

    J. Mater. Process. Technol.

    (2014)
  • T. Pottier et al.

    Inelastic heat fraction estimation from two successive mechanical and thermal analyses and full-field measurements

    Eur. J. Mech. A/Solids

    (2013)
  • B. Revil-Baudard et al.

    Plastic deformation of high-purity ??-titanium: model development and validation using the taylor cylinder impact test

    Mech. Mater.

    (2015)
  • D. Rittel et al.

    The dependence of the Taylor–Quinney coefficient on the dynamic loading mode

    J. Mech. Phys. Solids

    (2017)
  • D. Rittel

    On the conversion of plastic work to heat during high strain rate deformation of glassy polymers

    Mech. Mater.

    (1999)
  • M. Rosochowska et al.

    A new method of measuring thermal contact conductance

    J. Mater. Process. Technol.

    (2004)
  • A. Rusinek et al.

    Experiments on heat generated during plastic deformation and stored energy for trip steels

    Mater. Des.

    (2009)
  • V.M. Simões et al.

    The punch speed influence on warm forming and springback of two Al-Mg-Si alloys

    J. Manuf. Process.

    (2019)
  • A.T. Zehnder

    A model for the heating due to plastic work

    Mech. Res. Commun.

    (1991)
  • T. Zhang et al.

    Investigation on the plastic work-heat conversion coefficient of 7075-T651 aluminum alloy during an impact process based on infrared temperature measurement technology

    Acta Mech. Sin

    (2017)
  • Theory Manual, Version 6.11

    (2016)
  • Cited by (18)

    • Multi-stage hot deformation and dynamic recrystallization behavior of low-cost Ti-Al-V-Fe alloy via electron beam cold hearth melting

      2022, Journal of Materials Research and Technology
      Citation Excerpt :

      So, the lower recrystallization fraction and average grain size can be seen in Fig. 8(e). Generally, plastic deformations generate heat [39]. In Fig. 8(f), when the strain rate was increased to 1 s−1, although shorter deformation times resulted in more sub-strains, the heat of deformation generated by such high strain rates may lead to a sharp increase in the internal temperature of the material [40], allowing abnormal grain growth and resulting in an increase in average grain size.

    View all citing articles on Scopus
    View full text