Skip to main content
Log in

Inspection of ratcheting models for pathological error sensitivity and overparametrization

  • Published:
Meccanica Aims and scope Submit manuscript

Abstract

Accurate analysis of plastic strain accumulation under stress-controlled cyclic loading is vital for numerous engineering applications. Typically, models of plastic ratcheting are calibrated against available experimental data. Since actual experiments are not exactly accurate, one should check the identification protocols for pathological dependencies on experimental errors. In this paper, a step-by-step algorithm is presented to estimate the sensitivities of identified material parameters. As a part of the sensitivity analysis method, a new mechanics-based metric in the space of material parameters is proposed especially for ratcheting-related applications. The sensitivity of material parameters to experimental errors is estimated, based on this metric. Moreover, a relation between pathological error sensitivity and overparametrization is established. This relation gives rise to a new criterion of overparametrization. The advantages of the new overparametrization criterion are exposed and its plausibility is checked by alternative criteria, like the consideration of correlation matrices and validation of identified parameters on “unseen” data. For demonstration purposes, the accumulation of irreversible strain in the titanium alloy VT6 (Russian analog of Ti-6Al-4V) is analysed. Three types of phenomenological models of plastic ratcheting are considered. They are the Armstrong-Frederick model as well as the first and the second Ohno-Wang models. Based on real data, a new rule of isotropic hardening is proposed for greater accuracy of simulation. The ability of the sensitivity analysis to determine reliable and unreliable parameters is demonstrated.

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

Similar content being viewed by others

Notes

  1. Dealing with a general non-proportional loading, advanced models of directional distortional hardening are needed [12, 35, 42]. Moreover, simplified J2 yield conditions can be replaced by their anisotropic counterparts [2].

  2. Under stress-controlled cyclic loading, cyclic softening refers to a growth in cyclic strains under constant amplitude imposed cyclic stress.

  3. This was also shown in [22].

  4. Since the temperature changes are small in this study, we linearize the heat exchange term near \(\theta _0\). For example, the radiative heat transfer is proportional to \(\theta ^4- \theta _0^4 \approx 4 \theta _0^3 (\theta - \theta _0)\). As a result, the simple ansatz (29) is obtained.

  5. See also the discussion in [22].

  6. Recall that the Gauss-Newton method corresponds to the minimization of an auxiliary error functional, obtained by linearization of the model response.

References

  1. Abdel-Karim M, Ohno N (2000) Kinematic hardening model suitable for ratchetting with steady-state. Int J Plast 16:225–240

    MATH  Google Scholar 

  2. Adamus J, Lacki P (2016) Numerical simulation of forming titanium drawn part. Meccanica 51:391–400

    Google Scholar 

  3. Avril S, Grédiac M, Pierron F (2004) Sensitivity of the virtual field method to noisy data. Comput Mech 34:439–452

    MATH  Google Scholar 

  4. Bartel T, Osman M, Menzel A (2017) A phenomenological model for the simulation of functional fatigue in shape memory alloy wires. Meccanica 52(4–5):973–988

    MathSciNet  MATH  Google Scholar 

  5. Beck JV, Arnold KJ (2007) Parameter estimation in engineering and science. Wiley, New York

    MATH  Google Scholar 

  6. Benedetti KCB, Gonçalves PB, Silva FMA (2020) Nonlinear oscillations and bifurcations of a multistable truss and dynamic integrity assessment via a Monte Carlo approach. Meccanica 55:2623–2657

    MathSciNet  Google Scholar 

  7. Benedix U (2000) Parametrschätzung für elastisch-plastische Deformatiosgesetze bei Berücksichtigung lokaler und globaler Vergleichsgrößen. Dissertation. Chemnitz University

  8. Bratley P, Fox BL (1988) Algorithm 659: implementing Sobol’s quasirandom sequence generator. ACM Trans Math Softw 14(1):88–100

    MathSciNet  MATH  Google Scholar 

  9. Brünig M, Chyra O, Albrecht D, Driemeier L, Alves M (2008) A ductile damage criterion at various stress triaxialities. Int J Plast 24(10):1731–1755

    MATH  Google Scholar 

  10. Brun R, Reichert P (2001) Practical identifiability analysis of large environmental simulation models. Water Resour Res 37(4):1015–1030

    Google Scholar 

  11. Collins JA (1993) Failure of materials in mechanical design: analysis, prediction, prevention. Wiley, New York

    Google Scholar 

  12. François M (2001) A plasticity model with yield surface distortion for non proportional loading. Int J Plast 17:703–717

    MATH  Google Scholar 

  13. Grédiac M, Pierron F (2004) Applying the virtual fields method to the identification of elasto-plastic constitutive parameters. Int J Plast 22:602–627

    MATH  Google Scholar 

  14. Halphen G, Nguen QS (1975) Sur les matériaux standard généralisés. J M’ec 14:39–63

    MATH  Google Scholar 

  15. Harth T, Lehn J (2007) Identification of material parameters for inelastic constitutive models using stochastic methods. GAMM-Mitt 30(2):409–429

    MathSciNet  MATH  Google Scholar 

  16. Harth T, Schwan S, Lehn J, Kollmann FG (2004) Identification of material parameters for inelastic constitutive models: statistical analysis and design of experiments. Int J Plast 20:1403–1440

    MATH  Google Scholar 

  17. Hashiguchi K (1989) Subloading surface model in unconventional plasticity. Int J Solids Struct 25(8):917–945

    MATH  Google Scholar 

  18. Hashiguchi K (2017) Foundations of elastoplasticity: subloading surface model

  19. Hassan T, Kyriakides S (1994) Ratcheting of cyclically hardening and softening materials: I. Uniaxial behaviour. Int J Plast 10(2):149–184

    MATH  Google Scholar 

  20. Haupt P (2013) Continuum mechanics and theory of materials. Springer, Berlin

    Google Scholar 

  21. Kang G, Liu Y, Ding J, Gao Q (2009) Uniaxial ratcheting and fatigue failure of tempered 42crmo steel: damage evolution and damage-coupled visco-plastic constitutive model. Int J Plast 25(5):838–860

    MATH  Google Scholar 

  22. Kaygorodtseva AA, Kapustin VI, Zakharchenko KV, Shutov AV (2020) On the ratcheting of the vt6 alloy in a range of loading scenarios. J Phys 1666(2020):012020

    Google Scholar 

  23. Lemaitre J (1984) A three-dimensional ductile damage model applied to deep-drawing forming limits. Mech Behav Mater 2:1047–1053

    Google Scholar 

  24. Leuschner M (2018) Numerically efficient computational homogenization: Fourier-accelerated nodal solvers and reduced basis model order reduction. University of Stuttgart

  25. Lourakis MIA (2005) A brief description of the Levenberg–Marquardt algorithm implemented by Levmar. Found Res Technol 4(1):1

    Google Scholar 

  26. Lion A (2000) Constitutive modelling in finite thermoviscoplasticity: a physical approach based on nonlinear rheological elements. Int J Plast 16:469–494

    MATH  Google Scholar 

  27. Miehe C (2002) Strain-driven homogenization of inelastic microstructures and composites based on an incremental variational formulation. Int J Numer Methods Eng 55(11):1285–1322

    MathSciNet  MATH  Google Scholar 

  28. Nelder JA, Mead R (1965) A simplex method for function minimization. Comput J 7(4):308–313

    MathSciNet  MATH  Google Scholar 

  29. Niederreiter H (1978) Quasi-Monte Carlo methods and pseudo-random numbers. Bull Am Math Soc 84(6):957–1041

    MathSciNet  MATH  Google Scholar 

  30. Ohno N, Wang JD (1993) Kinematic hardening rules with critical state of dynamic recovery, Part I: formulation and basic features for ratchetting behavior. Int J Plast 9(3):375–390

    MATH  Google Scholar 

  31. Rao SS (2019) Engineering optimization: theory and practice. Wiley, New York

    Google Scholar 

  32. Salamon R, Kamiński H, Fritzkowski P (2020) Estimation of parameters of various damping models in planar motion of a pendulum. Meccanica 107:2411

    MathSciNet  Google Scholar 

  33. Seibert T, Lehn J, Schwan S, Collmann FG (2000) Identification of material parameters for inelastic constitutive models: stochastic simulations for the analysis of deviations. Continuum Mech Thermodyn 12:95–120

    MathSciNet  MATH  Google Scholar 

  34. Shutov AV, Ihlemann J (2011) On the simulation of plastic forming under consideration of thermal effects. Materialwissensch Werkstofftec 42(7):632–638

    Google Scholar 

  35. Shutov AV, Ihlemann J (2012) A viscoplasticity model with an enhanced control of the yield surface distortion. Int J Plast 39:152–167

    Google Scholar 

  36. Shutov AV, Silbermann CB, Ihlemann J (2015) Ductile damage model for metal forming simulations including refined description of void nucleation. Int J Plast 71:195–217

    Google Scholar 

  37. Shutov AV, Kaygorodtseva AA (2019) Parameter identification in elasto-plasticity: distance between parameters and impact of measurement errors. ZAMM 99:8

    MathSciNet  Google Scholar 

  38. Shutov AV, Kaygorodtseva AA (2020) Sample shapes for reliable parameter identification in elasto-plasticity. Acta Mech 231:4761

    Google Scholar 

  39. Shutov AV, Kreißig R (2008) Finite strain viscoplasticity with nonlinear kinematic hardening: phenomenological modeling and time integration. Comput Methods Appl Mech Eng 197:2015–2029

    MathSciNet  MATH  Google Scholar 

  40. Shutov AV, Kreißig R (2010) Regularized strategies for material parameter identification in the context of finite strain plasticity. Technisch Mech 30(1–3):280–295

    Google Scholar 

  41. Shutov AV, Larichkin AY, Shutov VA (2017) Modelling of cyclic creep in the finite strain range using a nested split of the deformation gradient. ZAMM 97(9):1083–1099

    MathSciNet  Google Scholar 

  42. Shutov AV, Panhans S, Kreißig R (2011) A phenomenological model of finite strain viscoplasticity with distortional hardening. ZAMM 91(8):653–680

    MathSciNet  MATH  Google Scholar 

  43. Sobol IM (1967) Distribution of points in a cube and approximate evaluation of integrals. Comput Math Math Phys 7:86–112

    MathSciNet  MATH  Google Scholar 

  44. Soize C (2017) Uncertainty quantification. Springer, Berlin

    MATH  Google Scholar 

  45. Surmiri A, Nayebi A, Rokhgireh H (2020) Application of anisotropic continuum damage mechanics in ratcheting characterization. Mech Adv Mater Struct. https://doi.org/10.1080/15376494.2020.1751353

    Article  Google Scholar 

  46. Ulloa J, Wambacq J, Alessi R, Degrande G, Francois S (2021) Phase-field modeling of fatigue coupled to cyclic plasticity in an energetic formulation. Comput Methods Appl Mech Eng 373:113473

    MathSciNet  MATH  Google Scholar 

  47. Vladimirov IN, Pietryga MP, Reese S (2008) On the modelling of non-linear kinematic hardening at finite strains with application to springback-comparison of time integration algorithms. Int J Numer Methods Eng 75(1):1–28

    MATH  Google Scholar 

  48. Wolszczak P, Lonkwic P, Cunha A Jr, Litak L, Molski S (2019) Robust optimization and uncertainty quantification in the nonlinear mechanics of an elevator brake system. Meccanica 54:1057–1069

    MathSciNet  Google Scholar 

  49. Wright S, Nocedal J (1999) Numerical optimization. Springer, Berlin

    MATH  Google Scholar 

  50. Yang X (2005) Low cycle fatigue and cyclic stress ratcheting failure behavior of carbon steel 45 under uniaxial cyclic loading. Int J Fatigue 27(9):1124–1132

    Google Scholar 

  51. Zhu SP, Lei Q, Wang QY (2017) Mean stress and ratcheting corrections in fatigue life prediction of metals. Fatigue Fract Eng Mater Struct 40(9):1343–1354

    Google Scholar 

Download references

Acknowledgements

The authors are thankful for stimulating discussions of the Monte Carlo method with Dr. I.N. Medvedev (Novosibirsk, Russia). We are also thankful to V.I. Kapustin and K.V. Zakharchenko (Novosibirsk, Russia) for providing experimental data on temperature evolution.

Funding

The research was supported by the State task project FWGG-2021-0012 2.3.1.3.1.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to A. V. Shutov.

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: Fast computation of \(\vec {p}^{\ (j)}\)

We discuss a quick computation of the parameter vectors \(\vec {p}^{\ (j)} \in \mathbb {R}^{n}\), corresponding to the jth draw of noisy data. The procedure is the same as in [38]. Recall that \(\overrightarrow{Exp} \in \mathbb {R}^{N_{\text {exp}}}\) is the vector of available experimental data, \(\overrightarrow{Mod}(\vec {p}) \in \mathbb {R}^{N_{\text {exp}}}\) is the corresponding modelling response, \(\vec {p} = (\vec {p}_c, \vec {p}_K) \in \mathbb {R}^{n}\) is the vector of unknown material parameters. Within the sensitivity analysis, the actual experimental data are replaced by the noisy data \(\overrightarrow{Exp} + \overrightarrow{Noise}\). The optimal set of parameters corresponding to noise-free data is denoted as \(\vec {p}^{\ *}\). The Jacobian of the model response at \(\vec {p}^{\ *}\) is the operator

$$\begin{aligned} \mathbf {J} := \frac{\partial \overrightarrow{Mod}(\vec {p})}{\partial \vec {p}}|_{\vec {p}^{\ *}} \in \mathbb {R}^{N_{\text {exp}} \times n}. \end{aligned}$$
(44)

Assuming only small changes in parameters, we linearize the model response near \(\vec {p}^{\ *}\):

$$\begin{aligned} \overrightarrow{Mod}^{lin}(\vec {p}) := \overrightarrow{Mod}(\vec {p}^{\ *}) + \mathbf {J}(\vec {p} - \vec {p}^{\ *}). \end{aligned}$$
(45)

Now, the parameter set \(\vec {p}^{\ (j)}\), \(j = 1, 2, ..., N_{\text {noise}}\) is the minimizer of the error function for noisy data

$$\begin{aligned}&\Phi ^{\text {noisy}}(\vec {p}) := \overrightarrow{Resid}^{\text {T}} \ \overrightarrow{Resid}, \end{aligned}$$
(46)
$$\begin{aligned}&\overrightarrow{Resid} := \overrightarrow{Exp} + \overrightarrow{Noise} - \overrightarrow{Mod}^{lin} \nonumber \\&\quad = \overrightarrow{Exp} + \overrightarrow{Noise} - \overrightarrow{Mod}(\vec {p}^{\ *}) - \mathbf {J}(\vec {p} - \vec {p}^{\ *}). \end{aligned}$$
(47)

Abbreviate by \(\overrightarrow{A}\) the following vector:

$$\begin{aligned} \overrightarrow{A} := \overrightarrow{\text {Exp}} + \overrightarrow{\text {Noise}} - \overrightarrow{\text {Mod}(\vec {p}^{\ *})} - \mathbf {J} \ \vec {p}^{\ *} . \end{aligned}$$
(48)

Then the error function is a quadratic form of \(\vec {p}\), given by

$$\begin{aligned} \Phi ^{\text {noisy}}(\vec {p}) = \big (\overrightarrow{A} - \mathbf {J} \vec {p}\big )^{\text {T}} \big (\overrightarrow{A} - \mathbf {J} \vec {p}\big ). \end{aligned}$$
(49)

Its derivative with respect to \(\vec {p}\) is a linear function of the unknown parameter vector \(\vec {p}\)

$$\begin{aligned} \frac{\displaystyle \partial \Phi ^{\text {noisy}}(\vec {p})}{\displaystyle \partial \vec {p} } = - 2 \big ( \overrightarrow{A} - \mathbf {J} \vec {p} \big )^{\text {T}} \ \mathbf {J}. \end{aligned}$$
(50)

The stationarity condition \(\frac{\displaystyle \partial \Phi ^{\text {noisy}}(\vec {p})}{\displaystyle \partial \vec {p} } = 0\) yields a system of linear algebraic equations with respect to \(\vec {p}\). Then the analytical solution is

$$\begin{aligned} \vec {p}^{\ (j)} = \big ( \mathbf {J}^{\text {T}} \mathbf {J} \big )^{-1} \big ( \mathbf {J} \big )^{\text {T}} \ \overrightarrow{A}. \end{aligned}$$
(51)

In fact, this semi-analytical solution represents a single iteration of the Gauss-Newton method [49].

Unfortunately, due to matrix multiplication, the condition number of \(\mathbf {J}^{\text {T}} \mathbf {J}\) can be very large. This effect may falsify the results of (51). To resolve this problem, \(\mathbf {Q} \mathbf {R}\) decomposition of \(\mathbf {J}\) should be implemented:

$$\begin{aligned}&\mathbf {J} = \mathbf {Q} \ \mathbf {R} \in \mathbb {R}^{N_{\text {exp}} \times n}, \quad \mathbf {Q} \in \mathbb {R}^{N_{\text {exp}} \times n},\nonumber \\&\quad \mathbf {Q}^{\text {T}} \mathbf {Q} = id_{\mathbb {R}^{n}} \in \mathbb {R}^{n \times n}, \quad \mathbf {R} \in \mathbb {R}^{n \times n}. \end{aligned}$$
(52)

Here, \(\mathbf {R}\) is upper triangular. Substituting this into (51), we have after some computations

$$\begin{aligned} \vec {p}^{\ (j)} = \mathbf {R}^{-1} \mathbf {Q}^{\text {T}} \ \overrightarrow{A}. \end{aligned}$$
(53)

Since the matrix \(\mathbf {R}^{-1} \mathbf {Q}^{\text {T}}\) is pre-computed, the parameter cloud is evaluated extremely efficiently even for a large number of draws (\(N_{\text {draws}} \ge 10,000\)). The condition number of the matrix \(\mathbf {R}\) behaves like the square root of the condition number of \(\mathbf {J}^{\text {T}} \mathbf {J}\). Thus, (53) gives more robust solution than (51).

Appendix B: Correlation matrices

Let \(\mathbf {J}\) be the Jacobian, defined in (44). The correlation matrix \(\mathbf {Corr} \in \mathbb {R}^{n \times n}\) is defined as follows [7]:

$$\begin{aligned} \mathbf {Corr}_{i j} = \mathbf {P}_{i j}/ \sqrt{ \mathbf {P}_{i i} \mathbf {P}_{j j} }, \quad \text {where} \quad \mathbf {P} = \big ( \mathbf {J}^{\text {T}} \ \mathbf {J} \big )^{-1} \in \mathbb {R}^{n \times n}. \end{aligned}$$
(54)

We say that there is a strong correlation between parameters \(p_i\) and \(p_j\), if \(\mathbf {Corr}_{i j} \approx \pm 1\). In that case, a slight change in \(p_i\) can be counteracted by a change in \(p_j\), still leaving the model response \(\overrightarrow{Mod}\) virtually the same. In such situations the minimum of the error functional \(\Phi\) lies in a “horizontal ravine” (cf. Fig. 10 in [32]). In the sense of Euclidean metric, a large correlation between parameters is characteristic for ill-defined optimization problems.

Table 7 Correlation matrix for AF model with \(N_{\text {branches}} = 2\)
Table 8 Correlation matrix for AF model with \(N_{\text {branches}} = 3\)
Table 9 Correlation matrix for AF model with \(N_{\text {branches}} = 4\)
Table 10 Correlation matrix for OW− I model with \(N_{\text {branches}} = 2\)
Table 11 Correlation matrix for OW-I model with \(N_{\text {branches}} = 3\)
Table 12 Correlation matrix for OW-I model with \(N_{\text {branches}} = 4\)
Table 13 Correlation matrix for OW-II model with \(N_{\text {branches}} = 2\)
Table 14 Correlation matrix for OW-II model with \(N_{\text {branches}} = 3\)
Table 15 Correlation matrix for OW-II model with \(N_{\text {branches}} = 4\)

Appendix C: Implementation of Sobol’s sequence

For each draw of the Monte Carlo method, the stochastic model (41) requires 40 independent random numbers \(\varepsilon _k \in \mathcal {N}(0,\sigma ^2)\) (20 numbers to obtain noisy data for each test). Within the quasi Monte Carlo method, they are obtained in the following way. First, we set the properties of the Sobol sequence. Dimensions is the number of terms of the Sobol sequence in each draw; we use \(Dimensions =40\). Skip is the number of initial points to omit from Sobol’s sequence, we put \(Skip = 10^3\). Leap is the interval between points of the sequence; \(Leap = 3 \cdot 10^2\) in our case.

Recall that \(N_{\text {noise}}\) is the number of draws. Calling Sobol’ generator [8] we obtain a matrix \(S \in \mathbb {R}^{N_{\text {noise}} \times Dimensions}\) of quasi-random numbers uniformly distributed over the interval [0, 1]. Then for the jth draw, the corresponding quasi-random variables with the normal distribution are:

$$\begin{aligned}&\varepsilon _{j,2i+1} = \cos ({2\pi } S_{j,2i+1}) \cdot \sqrt{-2 \ln {S_{j,2i+1}}}, \nonumber \\&\quad \varepsilon _{j,2i+2} = \sin ({2\pi } S_{j,2i+1}) \cdot \sqrt{-2 \ln {S_{j,2i+1}}}, \end{aligned}$$
(55)

for \(j = 1, 2, ..., N_{\text {noise}}\), \(i = 1, 2, ..., Dimensions\).

Appendix D: Fast computation of the distance

The fast computation of the distance between two sets of parameters relies on the linearization of the strain response function \(\varepsilon _{11}(t)\) with respect to the material parameters. For the fixed stress-controlled loading history (Fig. 15) we evaluate the derivative

$$\begin{aligned} d \mathbf {\varepsilon }/d \vec {p} (t) = \frac{\partial {\varepsilon _{1 1}(t,\vec {p})} }{\partial {\vec {p}}}|_{\vec {p} = \vec {p}^{\ *}}, \quad t \in [0,T_{\text {metric}}]. \end{aligned}$$
(56)

For the parameter set \(\vec {p}\) close to the center of the cloud \(\vec {p}^{\ *}\), the axial strain \(\varepsilon _{1 1}(t,\vec {p})\) is approximated as

$$\begin{aligned} \varepsilon _{1 1}(t,\vec {p}) = \varepsilon _{1 1}(t,\vec {p}^{\ *}) + d \mathbf {\varepsilon }/d \vec {p} (t) \cdot (\vec {p} - \vec {p}^{\ *}), \end{aligned}$$
(57)

where \(\varepsilon _{1 1}(t,\vec {p}^{\ *})\) is the strain history related to the center of the parameter cloud. Then the mechanics-based distance between \(\vec {p}^{\ *}\) and \(\vec {p}\) is

$$\begin{aligned} \text {dist}(\vec {p}, \vec {p}^{\ *}) = \max \limits _{t \in [0,T_{\text {metric}}]} | d \mathbf {\varepsilon }/d \vec {p} (t) \cdot (\vec {p} - \vec {p}^{\ *}) |. \end{aligned}$$
(58)

Appendix E: AF versus OW-II regarding error sensitivities

Table 5 says that the material parameters of the AF models are more sensitive to experimental errors than the parameters of the OW-II models. To obtain an intuitive insight into this effect, recall that for each instance of noisy data the set \(\vec {p}^{\ (j)}\) is computed through (48) and (53). Thus, the scatter of parameters \(\vec {p}^{\ (j)}\) depends on the matrix \(\mathbf {R} \in \mathbb {R}^{n \times n}\). Recall that \(\mathbf {R}\) is obtained by the \(\mathbf {Q} \mathbf {R}\) decomposition of the Jacobian \(\mathbf {J} \in \mathbb {R}^{N_{exp} \times n}\) (Appendix A). The condition numbers of \(\mathbf {R}\) are listed in Table 16 for both models; they are evaluated with respect to the \(l_2\) norm. Interestingly, the dependence of the condition number on \(N_{\text {branches}}\) is not monotonic. However, the AF models exhibit much larger condition number than the OW-II models in all the cases. Therefore, in the Euclidean \(l_2\) metric, the parameter vector of the AF models is more sensitive to the noise than for the OW-II models. This high sensitivity corresponds to the results shown in Table 5.

Table 16 Condition numbers of \(\mathbf {R}\)

As noted in Appendix A, the condition number of the matrix \(\mathbf {R}\) behaves like the square root of the condition number of \(\mathbf {J}^{\text {T}} \mathbf {J}\). The corresponding values are summarized in Table 17. Recall that the columns of \(\mathbf {J}\) are derivatives of the model response with respect to individual parameters. The large condition number of \(\mathbf {J}^{\text {T}} \mathbf {J}\) means that the columns of the Jacobian \(\mathbf {J}\) are “close” to being linearly dependent. Geometrically, this means that the columns “almost” lie in an \((n-1)\)-dimensional subspace of \(\mathbb {R}^{N_{exp}}\). Table 17 suggests that this is likely the case for the AF model.

Table 17 Condition numbers of \(\mathbf {J}^{\text {T}} \mathbf {J}\)

For the AF models, the minimum eigenvalues \(\lambda _{\text {min}}\) of the matrix \(\mathbf {R}\) are shown in Table 18. Since the eigenvalues are close to zero, the corresponding eigenvectors represent certain directions in the space of parameters, such that the model response \(\vec {Mod}(\vec {p})\) is almost the same along those directions (Table 18). Recalling (35), we conclude that a small increase in \(c_1\) can be compensated by an increase in \(\gamma\) and \(\beta\).

Table 18 Minimum eigenvalues and eigenvectors of \(\mathbf {R}\) for the AF models

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kaygorodtseva, A.A., Shutov, A.V. Inspection of ratcheting models for pathological error sensitivity and overparametrization. Meccanica 57, 1975–2000 (2022). https://doi.org/10.1007/s11012-022-01533-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11012-022-01533-5

Keywords

Navigation