Log in

Big Data Nanoindentation and Analytics Reveal the Multi-Staged, Progressively-Homogenized, Depth-Dependent Upscaling of Rocks’ Properties

  • Original Paper
  • Published:
Rock Mechanics and Rock Engineering Aims and scope Submit manuscript

Abstract

This paper presents a newly observed phenomenon of upscaling of rocks’ properties using big data nanoindentation and analytics involving Gaussian mixture modeling (GMM), leading to characterizing the cross-scale mechanical properties of four shales and one sandstone. A large number (i.e., ~ 500) of statistical indentation measurements to depths of 6–8 μm were performed on each rock, resulting in continuous depth-dependent hardness and Young’s modulus data from unknown phases, which were then segmented at various depths to extract an array of discretized subdatasets. Two-dimensional GMM of each subdataset yields the number, fraction, and properties of mechanically distinct phases, and re-assembly of these results leads to clearly discernible property-depth curves. Such improved data analytics consisting of data segmentation, GMM deconvolution, and re-assembly enables the transformation of a massive number of chaotic curves from unknown phases into a few discernible lines corresponding to identified phases, from which the mechanical properties of individual phases are accurately determined at relatively small depths. With increasing depth, initially unique mechanical properties of individual phases undergo multistage merging at the intermediate mesoscale and progressively homogenize into a unified value at large depths or macroscale (e.g., >  ~ 5 μm), which is regarded as the bulk rock’s properties. More importantly, such depth-dependent transition and progressive merging and homogenization actually manifest the micromechanics of nanoindentation on a heterogeneous composite, including the indentation surround effect and rock’s microstructure (e.g., sizes and spacings of different solid particles and their properties). Compared to different micromechanical upscaling models, this newly developed big data indentation technique and pertinent data analytics enable more accurate, multi-parameter, and cross-scale characterization of highly heterogeneous materials and explicitly uncover the multi-staged, progressively-homogenized, depth-dependent upscaling of elasticity from individual constituents at the nanoscale to merged virtual interface phases at the mesoscale and to bulk material at the macroscale.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Subscribe and save

Springer+ Basic
EUR 32.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or Ebook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

Price includes VAT (Canada)

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

Similar content being viewed by others

References

Download references

Acknowledgements

The authors are grateful to China National Petroleum Corporation and Solvay for providing the shale and sandstone samples, respectively. The first author acknowledges the Graduate Research Scholarship from the US Clay Minerals Society that partially supported this study. The authors also thank Dr. Marco Keiluweit and Ms. Mariela Garcia in the Stockbridge School of Agriculture at the University of Massachusetts Amherst for the TOC measurements. This work also made use of the Shared Experimental Facilities (i.e., X-ray diffractometer) at MIT, supported in part by the MRSEC Program of the National Science Foundation under award number DMR1419807.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Guo** Zhang.

Additional information

Publisher's Note

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

Appendices

Appendix A: Criteria for Results Filtering

For the purpose of clarification in the following discussion, use of different subscripts is explained here: (1) μm, n, r denotes the mean values of H and E of the mth phase in the nth solution derived from the model with r phases, where m = 1, 2, …, k, n = 1, 2, …, 1000, and r = 1, 2, …, N; (2) the same notation rule is applied to other parameters, except that logLn, r represents the logarithmic likelihood value of the nth solution of the model with r phases. Not all the 8000 fitted solutions were desired with the consideration of both the statistical perspective and the rationality in mechanical testing, so only those solutions that would satisfy all the following criteria (1)–(4) were selected for further analysis:

  1. 1.

    The successful application of statistical nanoindentation is based on an important premise that the mechanical properties of individual phases of the studied composites have sufficient contrast that can be captured by nanoindentation testing at small scales. It means that a rational solution should not display severe overlaps between the adjacent distributions of two or more different phases. For any two distributions (i.e., corresponding to Phase A and Phase B) in a solution with r > 1, the pairwise overlap between Phase A and B can be defined as ωab = ωa|b + ωb|a, where ωa|b given below is the misclassification probability that the random variable x originates from Phase B but is mistakenly assigned to Phase A, and vice versa (Maitra and Melnykov 2010):

    $$\omega_{{\text{a|b}}} = \Pr \left[ {\pi_{{\text{b, n, r }}} g({\mathbf{x}} \, ; \, {\varvec{\mu}}_{{\text{b, n, r}}} \, , \, {{\varvec{\Sigma}}}_{{\text{b, n, r}}} ) < \pi_{{\text{a, n, r }}} g({\mathbf{x}} \, ; \, {\varvec{\mu}}_{{\text{a, n, r}}} \, , \, {{\varvec{\Sigma}}}_{{\text{a, n, r}}} )|{\mathbf{x}}\sim g({\varvec{\mu}}_{{\text{b, n, r}}} \, , \, {{\varvec{\Sigma}}}_{{\text{b, n, r}}} )} \right].$$
    (10)

    In this study, the pairwise overlaps of all phases were calculated using the R package MixSim (Melnykov et al. 2012), and the maximum pairwise overlap was limited to 0.5, i.e. the solution that had two phases with over 50% chance of misclassification was excluded from subsequent analysis.

  2. 2.

    According to the data published in the literature (Zhang et al. 2013; Abedi et al. 2016; Akono and Kabir 2016; Sun et al. 2018), a positive correlation is usually observed between the hardness and Young’s modulus measured by nanoindentation. Therefore, the covariance term of each unrestricted variance–covariance matrix should be positive. Likewise, for any two distributions (e.g., Phase A and Phase B when r > 1) in a solution, their mean values μ = (μ_H, μ_E) were required to satisfy that both μ_Ea, n, r ≥ μ_Eb, n, r and μ_Ha, n, r ≥ μ_Hb, n, r held true.

  3. 3.

    For any two different solutions (e.g., Solutions C and D) having the same k (i.e., using the same model to fit the data but with different initializing parameters), they should have enough contrast in their estimates of parameters to be considered “unique”. In this work, the maximum difference calculated as follows was required to be greater than 5% if Solution C was not a duplicate of Solution D (presumably μ_Hm+1, n, r ≥ μ_Hm, n, r and μ_Em+1, n, r ≥ μ_Em, n, r):

    $$\max \left\{ \begin{gathered} \max \left( {\left| {\frac{{\pi_{{\text{1, c, r}}} - \pi_{{\text{1, d, r}}} }}{{\min \;(\pi_{{\text{1, c, r}}} , \, \pi_{{\text{1, d, r}}} )}}} \right| \, , \ldots , \, \left| {\frac{{\pi_{{\text{k, c, r}}} - \pi_{{\text{k, d, r}}} }}{{\min \;(\pi_{{\text{k, c, r}}} , \, \pi_{{\text{k, d, r}}} )}}} \right|} \right) \\ \, \max \left( {\left| {\frac{{\mu \_H_{{\text{1, c, r}}} - \mu \_H_{{\text{1, d, r}}} }}{{\min \;(\mu \_H_{{\text{1, c, r}}} , \, \mu \_H_{{\text{1, d, r}}} )}}} \right| \, , \ldots , \, \left| {\frac{{\mu \_H_{{\text{k, c, r}}} - \mu \_H_{{\text{k, d, r}}} }}{{\min \;(\mu \_H_{{\text{k, c, r}}} , \, \mu \_H_{{\text{k, d, r}}} )}}} \right|} \right) \, \\ \max \left( {\left| {\frac{{\mu \_E_{{\text{1, c, r}}} - \mu \_E_{{\text{1, d, r}}} }}{{\min \;(\mu \_E_{{\text{1, c, r}}} , \, \mu \_E_{{\text{1, d, r}}} )}}} \right| \, , \ldots , \, \left| {\frac{{\mu \_E_{{\text{k, c, r}}} - \mu \_E_{{\text{k, d, r}}} }}{{\min \;(\mu \_E_{{\text{k, c, r}}} , \, \mu \_E_{{\text{k, d, r}}} )}}} \right|} \right) \\ \end{gathered} \right\} \ge 0.05.$$
    (11)

    For any two solutions (e.g., Solutions E and F) having different k, the solution derived from a more complicated model (i.e., model with more parameters) should provide a higher logarithmic likelihood, i.e., logLe, r < logLf, r+1.

  4. 4.

    For Gaussian mixtures with unrestricted covariance matrices, it is almost inevitable to have some spurious solutions (Seo and Kim 2012). Generally, a spurious solution has one or more phases which overfit a small localized random pattern in the data rather than any other underlying group structure, resulting in a solution with a high likelihood but very small generalized variance and mixing proportion of the overfitted phase (McLachlan and Peel 2000; Kim and Seo 2014). In addition, prior studies on composites have shown that their mechanical properties can only be significantly enhanced when the amount of additive or filling phase reaches above ~ 5 wt.% (Maiti and Subbarao 1991; Koerner et al. 2004; Sreekanth et al. 2009). Thus, the final constraint specified a critical concentration of 5% to identify and eliminate the spurious solutions, i.e. the fraction of each phase for a qualified solution should be at least 5% (i.e., πm, n, r ≥ 0.05).

Appendix B: Upscaling Models for Calculation of Bulk Rock Properties

In geophysics, the Voigt and Reuss models are widely used to estimate the theoretical maximum and minimum elastic moduli of the dry rock frame, respectively, based on the known constituents and their volumetric fractions (Sone and Zoback 2013a, b). The Voigt upper bound is determined by assuming that the strain is uniform throughout the bulk rock, while the Reuss lower bound is obtained by assuming an iso-stress situation. For a heterogeneous rock with k phases, the two bounds are then calculated by:

$$E_{{{\text{Voigt}}}} = \sum\limits_{{r{ = 1}}}^{k} {f_{{\text{r}}} E_{{\text{r}}} } ,$$
(12)
$$\frac{1}{{E_{{{\text{Reuss}}}} }} = \sum\limits_{{r{ = 1}}}^{k} {\frac{{f_{{\text{r}}} }}{{E_{{\text{r}}} }}} ,$$
(13)

where k is the number of mechanically distinct phases; Er the Young’s modulus of the rth identified phase with a volumetric fraction of fr. Then the effective Young’s modulus of the bulk rock is often estimated by taking the average of the upper and lower bounds, known as the Voigt-Reuss-Hill (VRH) average:

$$E_{{{\text{VRH}}}} = \frac{{E_{{{\text{Voigt}}}} + E_{{{\text{Reuss}}}} }}{2}.$$
(14)

However, the stress and strain are generally nonuniform in rocks. Similar to the aforementioned bounding methods, the Mori–Tanaka (MT) model uses the average local stress and strain fields in the matrix-inclusion structure of a composite to predict the homogenized moduli, Khom and Ghom, of the composites (Mori and Tanaka 1973; Benveniste 1987), which can be determined by (da Silva et al. 2013; Huang et al. 2014):

$$K_{\hom } = \left( {\sum\limits_{{r{ = 1}}}^{k} {\frac{{f_{{\text{r}}} K_{{\text{r}}} }}{{3K_{{\text{r}}} + 4G_{1} }}} } \right)\left( {\sum\limits_{{r{ = 1}}}^{k} {\frac{{f_{{\text{r}}} }}{{3K_{{\text{r}}} + 4G_{1} }}} } \right)^{ - 1} ,$$
(15)
$$G_{\hom } = \frac{{\sum\nolimits_{{r{ = 1}}}^{k} {f_{{\text{r}}} G_{{\text{r}}} \left[ {G_{1} \left( {9K_{1} + 8G_{1} } \right) + 6G_{{\text{r}}} \left( {K_{1} + 2G_{1} } \right)} \right]^{ \, - 1} } }}{{\sum\nolimits_{{r{ = 1}}}^{k} {f_{{\text{r}}} \left[ {G_{1} \left( {9K_{1} + 8G_{1} } \right) + 6G_{{\text{r}}} \left( {K_{1} + 2G_{1} } \right)} \right]^{ \, - 1} } }},$$
(16)
$$E_{\hom } = \frac{{9K_{\hom } G_{\hom } }}{{3K_{\hom } + G_{\hom } }},$$
(17)

where K1 and G1 particularly denote the bulk and shear moduli of the matrix in the composite. The bulk modulus Kr and shear modulus Gr of the rth phase can be calculated by its Young’s modulus Er and Poisson’s ratio νr, which is assumed to be 0.2 for all phases:

$$K = \frac{E}{3(1 - 2\nu )},$$
(18)
$$G = \frac{E}{2(1 + \nu )}.$$
(19)

The third model used in this study is the self-consistent (SC) approximation that estimates the effective moduli of a composite by not only using the volumetric fractions and properties of individual phases, but also including their geometric information, such as sphere, needle, and disk (Budiansky 1965; Wu 1966; O’Connell and Budiansky 1974, 1977). The self-consistent (SC) effective moduli \(K_{{{\text{SC}}}}^{*}\) (i.e., bulk modulus) and \(G_{{{\text{SC}}}}^{*}\) (i.e., shear modulus) of the bulk composite can be approximated by (Berryman 1980, 1995):

$$\sum\limits_{{r{ = 1}}}^{k} {f_{{\text{r}}} (K_{{\text{r}}} - K_{{{\text{SC}}}}^{*} )P^{{*{\text{r}}}} } = 0,$$
(20)
$$\sum\limits_{{r{ = 1}}}^{k} {f_{{\text{r}}} (G_{{\text{r}}} - G_{{{\text{SC}}}}^{*} )Q^{{*{\text{r}}}} } = 0,$$
(21)

where P*r and Q*r are geometric factors of the rth phase (refer to Table 4 in Berryman 1995). Eqs. (20) and (21) are coupled and thus need to be solved simultaneously by iterations, and the self-consistent effective Young’s modulus can then be calculated by Eq. (17) with the converged solutions of \(K_{{{\text{SC}}}}^{*}\) and \(G_{{{\text{SC}}}}^{*}\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Luo, S., Kim, D., Wu, Y. et al. Big Data Nanoindentation and Analytics Reveal the Multi-Staged, Progressively-Homogenized, Depth-Dependent Upscaling of Rocks’ Properties. Rock Mech Rock Eng 54, 1501–1532 (2021). https://doi.org/10.1007/s00603-020-02337-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00603-020-02337-3

Keywords

Navigation