Skip to main content
Log in

Advantages of a 3-parameter reduced constitutive model for the measurement of polymers elastic modulus using tensile tests

  • Published:
Mechanics of Time-Dependent Materials Aims and scope Submit manuscript

Abstract

Exact measurements of the rheological parameters of time-dependent materials are crucial to improve our understanding of their intimate relation to the internal bulk microstructure. Concerning solid polymers and the apparently simple determination of Young’s modulus in tensile tests, international standards rely on basic protocols that are known to lead to erroneous values. This paper describes an approach allowing a correct measurement of the instantaneous elastic modulus of polymers by a tensile test. It is based on the use of an appropriate reduced model to describe the behavior of the material up to great strains, together with well-established principles of parameter estimation in engineering science. These principles are objective tools that are used to determine which parameters of a model can be correctly identified according to the informational content of a given data set. The assessment of the methodology and of the measurements is accomplished by comparing the results with those obtained from two other physical experiments, probing the material response at small temporal and length scales, namely, ultrasound measurements with excitation at 5 MHz and modulated nanoindentation tests over a few nanometers of amplitude.

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

Similar content being viewed by others

References

  • André, S., Meshaka, Y., Cunat, C.: Rheological constitutive equation of solids: a link between models based on irreversible thermodynamics and on fractional order derivative equations. Rheol. Acta 42, 500–515 (2003)

    Article  Google Scholar 

  • André, S., Renault, N., Meshaka, Y., Cunat, C.: From the thermodynamics of constitutive laws to thermomechanical experimental characterization of materials: an assessment based on inversion of thermal images. Contin. Mech. Thermodyn. 24, 1–20 (2012). doi:10.1007/s00161-011-0205-x

    Article  Google Scholar 

  • Arruda, E.M., Boyce, M.C.: A three-dimensional constitutive model for the large stretch behavior of rubber elastic materials. J. Mech. Phys. Solids 41–42, 389 (1993)

    Article  Google Scholar 

  • Aster, R.C., Borchers, B., Thurber, C.H.: Parameter Estimation and Inverse Problems, 2nd edn. Academic Press, San Diego (2013)

    MATH  Google Scholar 

  • Baravian, C., André, S., Renault, N., Moumini, N., Cunat, C.: Optical techniques for in situ dynamical investigation of plastic damage. Rheol. Acta 47, 555–564 (2008)

    Article  Google Scholar 

  • Bartczak, Z., Kozanecki, M.: Influence of molecular parameters on high-strain deformation of polyethylene in the plane-strain compression, Part I. Stress–strain behaviour. Polymer 46, 8210–8221 (2005)

    Article  Google Scholar 

  • Beck, J.V., Arnold, K.J.: Parameter Estimation in Engineering and Science. Wiley, New York (1977)

    MATH  Google Scholar 

  • Biot, M.A.: Linear thermodynamics and the mechanics of solids. In: Proceeding of the Third US National Congress of Applied Mechanics, pp. 1–18. ASME, New York (1958)

    Google Scholar 

  • Blaise, A., Baravian, C., André, S., Dillet, J., Michot, L.J., Mokso, R.: Investigation of the mesostructure of a mechanically deformed HDPE by synchrotron microtomography. Macromolecules 43, 8143–8152 (2010). doi:10.1021/ma101033b

    Article  Google Scholar 

  • Blaise, A., Baravian, C., Dillet, J., Michot, L.J., André, S.: Characterization of the mesostructure of HDPE under “in-situ” uniaxial tensile test by incoherent polarized steady-light transport. J. Polym. Sci., Part B, Polym. Phys. 50(5), 328–337 (2011). doi:10.1002/polb.23020

    Article  Google Scholar 

  • Callen, H.B.: Thermodynamics and an Introduction to Thermostatics. Wiley, New York (1985)

    MATH  Google Scholar 

  • Cooreman, S., Lecompte, D., Sol, H., Vantomme, J., Debruyne, D.: Elasto-plastic material parameter identification by inverse methods: calculation of the sensitivity matrix. Int. J. Solids Struct. 44(13), 4329–4341 (2007)

    Article  MATH  Google Scholar 

  • Cunat, C.: A thermodynamic theory of relaxation based on a distribution of non-linear processes. J. Non-Cryst. Solids 131(133), 196–199 (1991)

    Article  Google Scholar 

  • Cunat, C.: The DNLR approach and relaxation phenomena: Part I: Historical account and DNLR formalism. Mech. Time-Depend. Mater. 5, 39–65 (2001)

    Article  Google Scholar 

  • Faccio-Toussaint, E., Ayadi, Z., Pilvin, P., Cunat, C.: Modeling of the mechanical behavior of a nickel alloy by using a time-dependent thermodynamic approach to relaxations of continuous media. Mech. Time-Depend. Mater. 5, 1–25 (2001)

    Article  Google Scholar 

  • De Donder, T.: Thermodynamic Theory of Affinity: A Book of Principle. Oxford University Press, London (1936)

    Google Scholar 

  • Farge, L., André, S., Meneau, F., Dillet, J., Cunat, C.: A common multiscale feature of the deformation mechanisms of a semi-crystalline polymer. Macromolecules 46(24), 9659–9668 (2013a). doi:10.1021/ma4019747/

    Article  Google Scholar 

  • Farge, L., André, S., Pawlak, A., Baravian, C., Irvine, S.C., Philippe, A.M.: A study of the deformation-induced whitening phenomenon for cavitating and non-cavitating semi-crystalline polymers. J. Polym. Sci., Part B, Polym. Phys. 51(10), 826–841 (2013b). doi:10.1002/polb.23267

    Article  Google Scholar 

  • Gaucher-Miri, V., François, P., Séguéla, R.: On the mechanisms of initiation and propagation of plastic instability in polyethylene under tensile drawing. J. Polym. Sci., Part B, Polym. Phys. 34, 1113–1125 (1996)

    Article  Google Scholar 

  • Gutman, E.: Mechanochemistry of Materials. Cambridge International Science, Cambridge (1988)

    Google Scholar 

  • G’sell, C., Hiver, J.M., Dahoun, A.: Experimental characterization of deformation damage in solid polymers under tension, and its interrelation with necking. Int. J. Solids Struct. 39, 3857–3872 (2002)

    Article  Google Scholar 

  • Haward, R.N.: Strain hardening of thermoplastics. Macromolecules 26, 5860–5869 (1993)

    Article  Google Scholar 

  • Haward, R.N.: Strain hardening of high density polyethylene. J. Polym. Sci., Part B, Polym. Phys. 45, 1090–1099 (2007)

    Article  Google Scholar 

  • Hiss, R., Hobeika, S., Lynn, C., Strobl, G.: Network stretching, slip processes, and fragmentation of crystallites during uniaxial drawing of polyethylene and related copolymers. A comparative study. Macromolecules 32, 4390–4403 (1999)

    Article  Google Scholar 

  • Krempl, E.: Relaxation behaviour and modeling. Int. J. Plast. 17, 1419–1436 (2001)

    Article  MATH  Google Scholar 

  • Kuiken, G.D.C.: Thermodynamics of Irreversible Processes: Applications to Diffusion and Rheology. Wiley, New York (1994)

    Google Scholar 

  • Legros, N., Jen, C.-K., Ihara, I.: Ultrasonic evaluation and application of oriented polymer rods. Ultrasonics 37, 291–297 (1999)

    Article  Google Scholar 

  • Le Rouzic, J., Delobelle, P., Vairac, P., Cretin, B.: Comparison of three different scales techniques for the dynamic mechanical characterization of two polymers (PDMS and SU8). Eur. Phys. J. Appl. Phys. 48, 11201 (2009)

    Article  Google Scholar 

  • Maugin, G., Muschink, W.: Thermodynamics with internal variables. Part I: General concepts. J. Non-Equilib. Thermodyn. 19, 217 (1994)

    Google Scholar 

  • Meixner, J.Z.: Thermodynamik und Relaxationserscheinungen. Z. Naturforsch. 4a, 504–600 (1949)

    MATH  Google Scholar 

  • Mrabet, K., Rahouadj, R., Cunat, C.: An irreversible model for semicristalline polymers submitted to multisequence loading at large strain. Polym. Eng. Sci. 45(1), 42 (2005)

    Article  Google Scholar 

  • Negahban, M.: Experimentally evaluating the equilibrium stress in shear of glassy polycarbonate. J. Eng. Mater. Technol. 128, 537–542 (2006)

    Article  Google Scholar 

  • Nowick, A.S., Berry, B.S.: Anelastic Relaxation in Crystalline Solids. Academic Press, New York (1972)

    Google Scholar 

  • Oliver, W.C., Pharr, G.M.: An improved technique for determining hardness and elastic modulus using load and displacement sensing indentation experiments. J. Mater. Res. 7(6), 1563 (1992)

    Article  Google Scholar 

  • Piché, S.: Ultrasonic velocity measurement for the determination of density in polyethylene. Polym. Eng. Sci. 24, 1354–1358 (1984)

    Article  Google Scholar 

  • Ponthot, J.-P., Kleinermann, J.-P.: A cascade optimization methodology for automatic parameter identification and shape/process optimization in metal forming simulation. Comput. Methods Appl. Mech. Eng. 195(41–43), 5472–5508 (2006)

    Article  MATH  Google Scholar 

  • Prigogine, I., Defay, R.: Thermodynamique chimique conformément aux méthodes de Gibbs et De Donder, Tomes I–II. Gauthier-Villars, Paris (1944–1946)

  • Qasmi, M., Delobelle, P.: Influence of the average roughness Rms on the precision of the Young’s modulus and hardness determination using nanoindentation technique with a Berkovich indenter. Surf. Coat. Technol. 201, 1191 (2006)

    Article  Google Scholar 

  • Rahouadj, R., Ganghoffer, J.F., Cunat, C.: A thermodynamic approach with internal variables using Lagrange formalism. Part I: General framework. Mech. Res. Commun. 30(2), 109 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  • Treloar, L.R.G.: The Physics of Rubber Elasticity. Clarendon, Oxford (1975)

    MATH  Google Scholar 

  • Tschoegl, N.W.: The Phenomenological Theory of Linear Viscoelastic Behaviour, an Introduction. Springer, Berlin (1989)

    Book  MATH  Google Scholar 

  • Van Melick, H.G.H., Govaert, L.E., Meijer, H.E.H.: On the origin of strain hardening in glassy polymers. Polymer 44, 2493 (2003)

    Article  Google Scholar 

  • Walter, E., Pronzato, L.: Identification of Parametric Models from Experimental Data. Springer, Heidelberg (1997)

    MATH  Google Scholar 

  • Wang, M.C., Güth, E.: Statistical theory of networks of non-Gaussian flexible chains. J. Chem. Phys. 20(7), 1144 (1952)

    Article  MathSciNet  Google Scholar 

  • Ye, J., Andre, S., Farge, L.: Kinematic study of necking in a semi-crystalline polymer through 3D digital image correlation. Int. J. Solids Struct. 59, 58–72 (2015)

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to S. André.

Appendix:  Sensitivity analysis fundamentals

Appendix:  Sensitivity analysis fundamentals

Let \(y^{\mathrm{meas}} ( t )\) be the measured stress output of a system (our material specimen), and \(y ( \boldsymbol{\upbeta},t )\) the theoretical stress output of the behavioral model with a parameter vector \(\boldsymbol{\upbeta}\) of dimension \(n\), representing in our case the \(n\) constitutive material parameters. The output error \(e ( t )\), or the residuals, can be defined as

$$ e(t) = y^{\mathrm{meas}}(t) - y ( \boldsymbol{\upbeta},t ). $$
(A.1)

The role of an estimator is to minimize this output error. A least squares criterion can be used classically and is written as follows:

$$ E_{\mathrm{LS}} = \sum_{i = 1}^{m} \bigl( y^{\mathrm{meas}} ( t_{i} ) - y ( \boldsymbol{\upbeta},t_{i} ) \bigr)^{2} $$
(A.2)

where the sum is over the successive acquisition times \(t_{i}\), up to a total number of measurement data points \(m\).

The minimization of the criterion is done when its derivatives with respect to parameters \(\beta_{j}\) \((j = 1\ldots n)\) are null:

$$\begin{aligned} \forall j \in [1,n],\frac{\partial E_{\mathrm{LS}}}{\partial \beta_{j}} = 0 \quad \to\quad & \sum _{i = 1}^{m} \biggl( \frac{\partial y ( \boldsymbol{\upbeta},t_{i} )}{\partial \beta_{j}} \bigl[ y^{\mathrm{meas}} ( t_{i} ) - y ( \boldsymbol{\upbeta},t_{i} ) \bigr] \biggr) = 0 \\ \quad \to\quad & \sum_{i = 1}^{m} \bigl( X_{j} ( \boldsymbol{\upbeta},t_{i} ) \bigl[ y^{\mathrm{meas}} ( t_{i} ) - y ( \boldsymbol{\upbeta},t_{i} ) \bigr] \bigr) = 0. \end{aligned}$$
(A.3)

From this equation, the sensitivity coefficient vector component \(X_{j}\) associated to the parameter \(\beta_{j}\) can be clearly recognized as

$$ X_{j} ( \boldsymbol{\upbeta},t_{i} ) = \frac{\partial y(\boldsymbol{\upbeta},t_{i})}{\partial \beta_{j}}. $$
(A.4a)

Normalized sensitivity coefficients,

$$ \tilde{X}_{j} = \beta_{j}\frac{\partial y(\boldsymbol{\upbeta},t)}{\partial \beta_{j}}, $$
(A.4b)

are introduced to graphically check the level of sensitivity and possible correlations between parameters: different intervals of the time-independent variable \(t\) may be advantageously considered. These coefficients are dimensionally homogeneous to the signal itself, the stress in MPa.

Sensitivity coefficients express how much a model reacts to some small variation of the parameters. Sensitivity coefficients play a fundamental role in the conditioning of the inverse parameter estimation process and, as a consequence, in the errors made in the estimates of confidence bounds. It is obvious that major sensitivities are sought for, when designing an experiment for metrological purposes. In almost all cases, sensitivity coefficients are nonlinear functions of the parameters themselves when the model \(y\) is a nonlinear function of \(\beta_{j}\).

Switching to a matrix formulation and defining vectors \(\mathbf{Y}^{\mathrm{meas}}\) and \(\mathbf{Y}\) as

$$ \mathbf{Y}^{\mathrm{meas}} = \begin{bmatrix} y^{\mathrm{meas}} ( t_{1} )\\ y^{\mathrm{meas}} ( t_{2} ) \\ \ldots\\ y^{\mathrm{meas}} ( t_{m} ) \end{bmatrix} \quad \mbox{and} \quad \mathbf{Y} = \begin{bmatrix} y ( \boldsymbol{\upbeta},t_{1} )\\ y ( \boldsymbol{\upbeta},t_{2} ) \\ \ldots \\ y ( \boldsymbol{\upbeta},t_{m} ) \end{bmatrix} $$
(A.5)

makes the minimization process expressed as

$$ \mathbf{X}^{T} \bigl( \mathbf{Y}^{\mathrm{meas}} - \mathbf{Y} \bigr) = 0 $$
(A.6)

with \(\mathbf{X}\), the \(m \times n\) rectangular sensitivity matrix, where \(m\) is the number of rows (dimension of the experimental observation time vector \(\mathbf{t}\)) and \(n\) is the number of columns (dimension of the parameter vector \(\boldsymbol{\upbeta}\)).

In the case of a linear model with respect to the parameters (linear estimation problem) for which the matrix of the sensitivity coefficients does not depend on the parameters, we have

$$ \mathbf{Y} = \mathbf{X}\boldsymbol{\upbeta}. $$
(A.7)

The estimated parameter vector, denoted by \(\hat{\boldsymbol{\upbeta}}\), corresponds to the value reached by \(\boldsymbol{\upbeta}\) when the criterion is minimized. Therefore, using (A.7), we can rewrite Eq. (A.6) as

$$ \mathbf{X}^{T} \bigl( \mathbf{Y}^{\mathrm{meas}} - \mathbf{X}\hat{\boldsymbol{\upbeta}} \bigr) = \boldsymbol{0}. $$
(A.8)

Relation (A.8) can be inverted to obtain the expression of \(\hat{\boldsymbol{\upbeta}}\) in the case of a linear estimation problem:

$$ \hat{\boldsymbol{\upbeta}} = \bigl( \mathbf{X}^{T} \mathbf{X} \bigr)^{ - 1} \mathbf{X}^{T} \mathbf{Y}^{\mathrm{meas}}. $$
(A.9)

However, most parameter estimation problems are not linear and require an iterative linearizing procedure. This is obtained by developing the solution at rank \(n+1\) in the neighborhood of the solution obtained for the prior iteration at rank \(n\):

$$ \mathbf{Y}^{(n + 1)} = \mathbf{Y}^{(n)} + \mathbf{X}^{(n)} \bigl( \hat{\boldsymbol{\upbeta}}^{(n + 1)} - \hat{\boldsymbol{\upbeta}}^{(n)} \bigr). $$
(A.10)

If relation (A.6) is written at rank \(n+1\) for parameters estimated at rank \(n\), we obtain \(\mathbf{X}^{T(n)} ( \mathbf{Y}^{\mathrm{meas}} - \mathbf{Y}^{(n + 1)} ) = \boldsymbol{0}\), and when this is combined with relation (A.10), we obtain the following relation of recurrence between estimated parameters at rank \(n+1\) and rank \(n\):

$$ \hat{\boldsymbol{\upbeta}}^{(n + 1)} = \hat{\boldsymbol{\upbeta}}^{(n)} + \bigl( \mathbf{X}^{T(n)} \ \mathbf{X}^{(n)} \bigr)^{ - 1} \mathbf{X}^{T(n)} \bigl( \mathbf{Y}^{\mathrm{meas}} - \mathbf{Y}^{(n)} \bigr) $$
(A.11)

which defines the iterative procedure that can be used to estimate the parameters (Gauss–Newton algorithm).

The statistical estimator’s properties depend on the noise \(\varepsilon (t)\)of the signal. If the theoretical model is assumed to be unbiased (perfect) then we have

$$ Y^{\mathrm{meas}}(t) = Y(t,\boldsymbol{\upbeta} ) + \varepsilon (t). $$
(A.12)

If classical statistical assumptions are made concerning the experimental noise \(\varepsilon (t)\) in the measured stress signal (Beck and Arnold 1977), it is possible to obtain an approximation of the errors that can be made in the estimation process for the different parameters. These assumptions are:

  1. (a)

    Zero mean value of the signal in the absence of excitation, which corresponds to a zero expectancy for noise (expected value \(E(\boldsymbol{\upvarepsilon} ) = 0\));

  2. (b)

    Constant variance or standard deviation of the noise, \(V(\boldsymbol{\upvarepsilon} ) = \sigma_{0}^{ 2}\).

In the case of a 1st order linearized estimation, its expected value can be proved to be

$$ E(\hat{\boldsymbol{\upbeta}} ) = \boldsymbol{\upbeta}. $$
(A.13)

This means that there is no error or bias made on the identified parameters.

The variance–covariance matrix \(\boldsymbol{\Delta}\) of the estimated parameters (generalization of the scalar-valued variance to a higher dimension) involves the sensitivity coefficients. It is calculated as \(\boldsymbol{\Delta} = E [ ( \hat{\boldsymbol{\upbeta}} - E ( \hat{\boldsymbol{\upbeta}} ) ) ( \hat{\boldsymbol{\upbeta}} - E ( \hat{\boldsymbol{\upbeta}} ) )^{T} ] = \sigma_{0}^{2} ( \mathbf{X}^{T} \mathbf{X} )^{ - 1}\) which in the expanded form gives

$$ \boldsymbol{\Delta} = \begin{bmatrix} V(\beta_{1}) & {\mathop{\mathrm{cov}}} (\beta_{1},\beta_{2}) &\ldots & \mathrm{cov} (\beta_{1},\beta_{p})\\ {\mathop{\mathrm{cov}}} (\beta_{1},\beta_{2}) & V(\beta_{2}) &\ldots & \mathrm{cov} (\beta_{2},\beta_{p}) \\ \ldots &\ldots &\ldots &\ldots \\ {\mathop{\mathrm{cov}}} (\beta_{1},\beta_{p}) & {\mathop{\mathrm{cov}}} (\beta_{2},\beta_{p}) &\ldots & V(\beta_{p}) \end{bmatrix}. $$
(A.14)

A stochastic analysis can be made that consists of calculating the variance–covariance matrix through Eq. (A.14) theoretically according to a given noise and set of parameters. This symmetric square matrix has dimension equal to the number of parameters. The diagonal terms correspond directly to the variance of each parameter \(V(\beta_{j})\). They can be used to determine the error made for each parameter. This error can be expressed as a percentage

$$ \mathit{Err}(\beta_{j}) = \frac{\sqrt{V(\beta_{j})}}{\beta_{j}}. $$
(A.15)

The off-diagonal terms can be used to calculate the correlation coefficients \(\rho_{mn}\) which express the degree of correlation of the parameters,

$$ \rho_{rs} = \frac{\mathrm{cov} (\beta_{r},\beta_{s})}{\sqrt{V(\beta_{r})} \sqrt{V(\beta_{s})}}. $$
(A.16)

The \(\vert \rho_{rs} \vert \) values lie between 0 and 1. In the case of a model with strongly correlated parameters, the correlation coefficients are close to 1, which means that two columns of the sensitivity matrix \(X\) are nearly proportional to each other. The resulting confidence bounds interval for two correlated parameters are therefore generally very high. This means that a large number of solutions exist for these two parameters to allow for a good fit to the experimental curve. A deterministic algorithm like the steepest gradient technique, used for the minimization process, is consequently very sensitive to the initial estimate made for the parameters. A strategy to produce first approximations using the physical background is highly recommended. But even so, the estimation problem is ill-conditioned and indicates to the experimentalist that the physical description involved is probably not appropriate and must be changed.

In the following considerations, the identifiability of the model parameters will be analyzed through the \(\tilde{\boldsymbol{\Delta}}\) array, which combines the variance–covariance matrix and the correlation matrix. \(\tilde{\boldsymbol{\Delta}}\) puts the main diagonal of \(\boldsymbol{\Delta}\) in place and the correlation coefficients on the off-diagonal terms,

$$ \tilde{\boldsymbol{\Delta}} = \begin{bmatrix} \mathit{Err}(\beta_{1}) & \rho_{12} &\ldots & \rho_{1p}\\ \rho_{12} & \mathit{Err}(\beta_{2}) &\ldots & \rho_{2p} \\ \ldots &\ldots &\ldots &\ldots \\ \rho_{1p} & \rho_{2p} &\ldots & \mathit{Err}(\beta_{p}) \end{bmatrix}. $$
(A.17)

When there is no model bias, or perfect agreement between the conditions of the experiment and the model based on it, the relation between the estimated parameter vector for the nonlinear estimation problem and its ‘exact’ value can be given at convergence by the following formula:

$$ \hat{\boldsymbol{\upbeta}} = \boldsymbol{\upbeta} + \bigl( \mathbf{X}^{T} \mathbf{X} \bigr)^{ - 1} \mathbf{X}^{T} \boldsymbol{\upvarepsilon}. $$
(A.18)

In this case, the residuals \(e ( t )\) that are the difference between the model and the data correspond exactly to the noise. Their standard deviation corresponds to the experimental standard deviation of the noise, and the residuals remain unsigned with no large fluctuation around the zero level.

The matrix \(\tilde{\boldsymbol{\Delta}}\) is a good tool to investigate the identification conditions of a parameter estimation problem.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Blaise, A., André, S., Delobelle, P. et al. Advantages of a 3-parameter reduced constitutive model for the measurement of polymers elastic modulus using tensile tests. Mech Time-Depend Mater 20, 553–577 (2016). https://doi.org/10.1007/s11043-016-9312-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11043-016-9312-1

Keywords

Navigation