Skip to main content
Log in

Inversion and Geodiversity: Searching Model Space for the Answers

  • Special Issue
  • Published:
Mathematical Geosciences Aims and scope Submit manuscript

Abstract

Geophysical inversion employs various methods to minimize the misfit between geophysical datasets and three-dimensional petrophysical distributions. Inversion techniques rely on many subjective inputs to provide a solution to a non-unique underdetermined problem, including the use of a priori model elements (i.e. a contiguous volume of the same litho-stratigraphic package), the a priori input model itself or inversion constraints. In some cases, inversion may produce a result that perfectly matches the observed geophysical data, but can still misrepresent the geological system. A workflow is presented here that offers objective methods to provide inputs to inversion: (1) simulations are performed to create a model suite that contains a range of geologically possible models; (2) stratigraphic variability is determined via uncertainty analysis to identify low certainty model regions and elements; (3) geodiversity analysis is then conducted to determine geometrical and geophysical extremes and commonalities within the model space; (4) geodiversity metrics are simultaneously analysed using principal component analysis to identify the contribution of different model elements toward overall model suite uncertainty; (5) principal component analysis also determines which models exhibit diverse or common geological and geophysical characteristics which (6) facilitate the selection of models as inputs to geophysical inversion. This workflow is applied to a three-dimensional model of the Ashanti Greenstone Belt, southwestern Ghana in West Africa in order to reduce the subjectivity incurred during decision making, explore the range of geologically possible models and provide geological constraints to the inversion process to produce geologically and geophysically robust suites of models. Results further suggest that three-dimensional uncertainty grids can optimize inversion processes and assist in finding geologically reasonable solutions.

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
Fig. 16
Fig. 17
Fig. 18

Similar content being viewed by others

References

  • Adadey K, Clarke B, Théveniaut H, Urien P, Delor C, Roig JY et al (2009) Geological map explanation—map sheet 0503 B (1:100 000), CGS/BRGM/Geoman, Geological Survey Department of Ghana (GSD). No. MSSP/2005/GSD/5a

  • Aitken ARA, Betts PG (2009) Multi-scale integrated structural and aeromagnetic analysis to guide tectonic models: an example from the eastern Musgrave Province, Central Australia. Tectonophysics 476: 418–435

  • Aitken ARA, Salmon ML, Kennett BLN (2013) Australia’s Moho: a test of the usefulness of gravity modelling for the determination of Moho depth. Tectonophysics 609:468–479

    Article  Google Scholar 

  • Batty M (1974) Spatial Entropy. Geogr Anal 6:1–31

    Article  Google Scholar 

  • Bertoncello A, Sun T, Li H, Mariethoz G, Caers J (2013) Conditioning surface-based geological models to well and thickness data. Math Geosci 45:873–893

    Article  Google Scholar 

  • Besl PJ, Jain RC (1986) Invariant surface characteristics for 3D object recognition in range images. Comput Vis Graph Image Process 33:33–80

    Article  Google Scholar 

  • Boschetti F, Moresi L (2001) Interactive inversion in geosciences. Geophysics 66:1226–34

    Article  Google Scholar 

  • Bremner D, Demaine E, Erickson J, Iacono J, Langerman S, Morin P et al (2005) Output-sensitive algorithms for computing nearest-neighbour decision boundaries. Discret Comput Geom 33:593–604

    Article  Google Scholar 

  • Calcagno P, Chilès JP, Courrioux G, Guillen A (2008) Geological modelling from field data and geological knowledge: part I. Modelling method coupling 3D potential-field interpolation and geological rules. Phys Earth Planet Inter 171:147–157

    Article  Google Scholar 

  • Caratori Tontini F, Cocchi L, Carmisciano C (2009) Rapid 3-D forward model of potential fields with application to the Palinuro Seamount magnetic anomaly (southern Tyrrhenian Sea, Italy). J Geophys Res 114:1978–2012

    Google Scholar 

  • Caumon G (2010) Towards stochastic time-varying geological modeling. Math Geosci 42:555–569

    Article  Google Scholar 

  • Clark DA (1997) Magnetic petrophysics and magnetic petrology: aids to geological interpretation of magnetic surveys. ASGO J Aust Geol Geophys 17:83–103

    Google Scholar 

  • Cooley M (2007) From judgment to calculation. AI Soc 21:395–409

    Article  Google Scholar 

  • Eisenlohr BN, Hirdes W (1992) The structural development of the early Proterozoic Birimian and Tarkwaian rocks of southwest Ghana, West Africa. J Afr Earth Sci 14:313–325

    Article  Google Scholar 

  • Feybesse J-L, Billa M, Guerrot C, Duguey E, Lescuyer J-L, Milesi J-P et al (2006) The paleoproterozoic Ghanaian province: geodynamic model and ore controls, including regional stress modeling. Precambrian Res 149:149–196

    Article  Google Scholar 

  • Friedman JH, Bentley J, Finkel RA (1977) An algorithm for finding best matches in logarithmic expected time. ACM Trans Math Softw 3:209–226

    Article  Google Scholar 

  • Frodeman R (1995) Geological reasoning: geology as an interpretive and historical science. Geol Soc Am Bull 107:960–968

    Article  Google Scholar 

  • Fullagar PK (2009) VPmg user documentation, version 5.2. Fullagar Geophysics, Report FGR01F-4: p 62

  • Fullagar PK, Hughes NA, Paine J (2000) Drilling-constrained 3D gravity interpretation. Explor Geophys 31:017–23

    Article  Google Scholar 

  • Fullagar PK, Pears G, Hutton D, Thompson A (2004) 3D gravity and aeromagnetic inversion for MVT lead-zinc exploration at Pillara, Western Australia. Explor Geophys 35:142–146

    Article  Google Scholar 

  • Fullagar PK, Pears GA, McMonnies B (2008) Constrained inversion of geologic surfaces—pushing the boundaries. Lead Edge 27:98–105

    Article  Google Scholar 

  • Gallardo LA, Meju MA (2011) Structure-coupled multiphysics imaging in geophysical sciences. Rev Geophys 49, RG1003

  • Gonzalez RC, Woods RE, Eddins SL (2003) Digital image processing using MATLAB. Prentice Hall, New Jersey

  • Guillen A, Calcagno P, Courrioux G, Joly A, Ledru P (2008) Geological modelling from field data and geological knowledge: part II. Modelling validation using gravity and magnetic data inversion. Phys Earth Planet Inter 171:158–169

    Article  Google Scholar 

  • Hotelling H (1931) The generalization of student’s ratio. Ann Math Stat 2:360–378

    Article  Google Scholar 

  • Huttenlocher DP, Klanderman GA, Rucklidge WJ (1993) Comparing images using the Hausdorff distance. IEEE Trans Pattern Anal Mach Intell 15:850–863

    Article  Google Scholar 

  • Jessell M (2001) Three-dimensional geological modelling of potential-field data. Comput Geosci 27: 455–465

  • Jessell MW, Ailleres L, de Kemp EA (2010) Towards an integrated inversion of geoscientific data: what price of geology? Tectonophysics 490:294–306

    Article  Google Scholar 

  • Jolliffe IT (2002) Principal component analysis, 2nd edn. Springer, New York

    Google Scholar 

  • Junner N (1940) Geology of the Gold Coast and Western Togoland with revised geological map. Memoir 11:40

    Google Scholar 

  • Kitson SAE (1928) Provisional geological map of the Gold Coast with Western Togoland with brief descriptive notes thereon, Gold Coast Geological Survey. Bulletin 2, Benham & Company Limited

  • Krzanowski WJ (1996) Principles of multivariate analysis: a user’s perspective. Oxford University Press, New York

    Google Scholar 

  • Lajaunie C, Courrioux G, Manuel L (1997) Foliation fields and 3D cartography in geology: principles of a method based on potential interpolation. Math Geol 29:571–584

    Article  Google Scholar 

  • Lane R, Guillen A (2005) Geologically-inspired Constraints for a Potential Field Litho-inversion Scheme. Proc IAMG’05 GIS Spat Anal 1: 181–6

  • Lelièvre PG, Oldenburg DW, Williams NC (2009) Integrating geological and geophysical data through advanced constrained inversions*. Explor Geophys 40:334–341

    Article  Google Scholar 

  • Lindsay MD, Aillères L, Jessell MW, de Kemp EA, Betts PG (2012) Locating and quantifying geological uncertainty in three-dimensional models: analysis of the Gippsland Basin, southeastern Australia. Tectonophysics 546–547:10–27

    Article  Google Scholar 

  • Lindsay MD, Jessell MW, Ailleres L, Perrouty S, de Kemp E, Betts PG (2013a) Geodiversity: exploration of 3D geological model space. Tectonophysics 594:27–37

    Article  Google Scholar 

  • Lindsay MD, Jessell MW, Perrouty S, Ailleres L, de Kemp EA, Betts PG (2013b) Making the link between geological and geophysical uncertainty: geodiversity in the Ashanti Greenstone Belt. Geophys J Int 195:903–922

    Article  Google Scholar 

  • Lisle RJ, Robinson JM (1995) The Mohr circle for curvature and its application to fold description. J Struct Geol 17:739–750

    Article  Google Scholar 

  • Lisle RJ, Toimil NC (2007) Defining folds on three-dimensional surfaces. Geology 35:519–22

    Article  Google Scholar 

  • Loh G, Hirdes W, Anani C, Davis DW, Vetter UK (1999) Explanatory notes for the geological map of Southwest Ghana 1: 100,000, Geologisches Jahrbuch, Reihe B, Heft 93\(,\) 150

  • Manger GE (1963) Porosity and bulk density of sedimentary rocks. US Geological Survey Bull. 1144-E

  • Mann JC (1993) Uncertainty in geology. In: Computers in geology—25 years of progress. Oxford University Press, New York, pp 241–54

  • Martelet G, Calcagno P, Gumiaux C, Truffert C, Bitri A, Gapais D et al (2004) Integrated 3D geophysical and geological modelling of the Hercynian Suture Zone in the Champtoceaux area (south Brittany, France). Tectonophysics 382:117–28

    Article  Google Scholar 

  • Martin R, Monteiller V, Komatitsch D, Perrouty S, Jessell M, Bonvalot S et al (2013) Gravity inversion using wavelet-based compression on parallel hybrid cpu/gpu systems: application to southwest ghana. Geophys J Int 195:1594–1619

    Article  Google Scholar 

  • Maxelon M, Renard P, Courrioux G, Brändli M, Mancktelow N (2009) A workflow to facilitate three-dimensional geometrical modelling of complex poly-deformed geological units. Comput Geosci 35: 644–658

  • Metelka V, Baratoux L, Naba S, Jessell MW (2011) A geophysically constrained litho-structural analysis of the Eburnean greenstone belts and associated granitoid domains, Burkina Faso, West Africa. Precambrian Res 190:48–69

    Article  Google Scholar 

  • Muxworthy AR, McClelland E (2000) Review of the low-temperature magnetic properties of magnetite from a rock magnetic perspective. Geophys J Int 140:101–114

    Article  Google Scholar 

  • Mynatt I, Bergbauer S, Pollard DD (2007) Using differential geometry to describe 3-D folds. J Struct Geol 29:1256–1266

    Article  Google Scholar 

  • O’Gorman L, Sammon MJ, Seul M (2008) Practical algorithms for image analysis: description, examples, programs, and projects, 2nd edn. Cambridge University Press, New York

  • Oberthür T, Vetter U, Davis DW, Amanor J (1998) Age constraints on gold mineralization and Palaeoproterozoic crustal evolution in the Ashanti belt of southern Ghana. Precambrian Res 89:129–143

    Article  Google Scholar 

  • Oldenburg DW (1974) Inversion and interpretation of gravity anomalies. Geophysics 39:526–36

    Article  Google Scholar 

  • Oldenburg DW, Li Y, Ellis RG (1997) Inversion of geophysical data over a copper gold porphyry deposit: a case history for Mt. Milligan Geophys 62:1419–1431

    Article  Google Scholar 

  • Olson CF, Huttenlocher DP (1997) Automatic target recognition by matching oriented edge pixels. IEEE Trans Image Process 6:103–113

    Article  Google Scholar 

  • Perrouty S (2012) Chapter 4–3D modelling of the Ashanti Belt, southwest Ghana, and lithostratigraphic control on gold occurences within the Sefwi Group. In: Evolution Structurale de la Ceinture Minéralisée d’Ashanti, Sud-Ouest Ghana. Unpublished thesis, Université Paul Sabatier (Toulouse III)

  • Perrouty S, Aillères L, Jessell MW, Baratoux L, Bourassa Y, Crawford B (2012) Revised Eburnean geodynamic evolution of the gold-rich southern Ashanti Belt, Ghana, with new field and geophysical evidence of pre-Tarkwaian deformations. Precambrian Res 204–205:12–39

    Article  Google Scholar 

  • Pigios J-P, Groves DI, Fletcher IR, McNaughton NJ, Snee LW (2003) Age constraints on Tarkwaian palaeoplacer and lode-gold formation in the Tarkwa-Damang district, SW Ghana. Mineralium Deposita 38: 695–714

  • Polanyi M (1962) Tacit knowing: its bearing on some problems of philosophy. Rev Modern Phys 34:601–616

    Article  Google Scholar 

  • Royse KR (2010) Combining numerical and cognitive 3D modelling approaches in order to determine the structure of the Chalk in the London Basin. Comput Geosci 36:500–511

    Article  Google Scholar 

  • Rucklidge WJ (1997) Efficiently locating objects using the Hausdorff distance. Int J Comput Vis 24:251–270

    Article  Google Scholar 

  • Saltus RW, Blakely RJ (2011) Unique geologic insights from “non-unique” gravity and magnetic interpretation. GSA Today 21:8

    Article  Google Scholar 

  • Scheidt C, Caers J (2009) Representing spatial uncertainty using distances and kernels. Math Geosci 41: 397–419

  • Schreiber D, Lardeaux J-M, Martelet G, Courrioux G, Guillen A (2010) 3-D modelling of Alpine Mohos in Southwestern Alps. Geophys J Int 180:961–975

    Article  Google Scholar 

  • Shannon CE (1948) A mathematical theory of communication. Bell Syst Tech J 27:379–423

    Article  Google Scholar 

  • Sim D-G, Kwon O-K, Park R-H (1999) Object matching algorithms using robust Hausdorff distance measures. IEEE Trans Image Process 8:425–429

    Article  Google Scholar 

  • Talwani M, Heirtzler JR (1964) In: Parks GA (ed) Computation of magnetic anomalies caused by two-dimensional bodies of arbitrary shape. Computers in the Mineral Industries. School of Earth Sciences, Stanford University, pp 464–480

  • Talwani M, Worzel JI, Landisman M (1959) Rapid gravity computations for twodimensional bodies with application to the Mendicino submarine fracture zone. J Geophys Res 64:49–59

    Article  Google Scholar 

  • Tarantola A (1984) Inversion of seismic relfection data in the acoustic approximation. Geophysics 49: 1259–1266

  • Tarantola A, Valette B (1982a) Generalized nonlinear inverse problems solved using the least squares criterion. Rev Geophys Space Phys 20:219–232

    Article  Google Scholar 

  • Tarantola A, Valette B (1982b) Inverse problems = quest for information. J Geophys: Zeitschrift fur Geophysik 50:159–170

  • Thore P, Shtuka A, Lecour M, Ait-Ettajer T, Cognot R (2002) Structural uncertainties: determination, management, and applications. Geophysics 67:840–852

    Article  Google Scholar 

  • Torvela T, Bond CE (2010) Do experts use idealised structural models? Insights from a deepwater fold-thrust belt. J Struct Geol 33:51–58

    Article  Google Scholar 

  • Vouillamoz N, Sue C, Champagnac JD, Calcagno P (2012) 3D cartographic modeling of the Alpine arc. Tectonophysics 579:131–143

    Article  Google Scholar 

  • Wang L, Suter D (2007) Learning and matching of dynamic shape manifolds for human action recognition. IEEE Trans Image Process 16:1646–1661

    Article  Google Scholar 

  • Wellmann JF, Horowitz FG, Schill E, Regenauer-Lieb K (2010) Towards incorporating uncertainty of structural data in 3D geological inversion. Tectonophysics 490:141–151

    Article  Google Scholar 

  • Wellmann JF, Regenauer-Lieb K (2011) Uncertainties have a meaning: information entropy as a quality measure for 3-D geological models. Tectonophysics 526–529:207–216

    Google Scholar 

  • Williams HA, Betts PG, Ailleres L (2009) Constrained 3D modeling of the mesoproterozoic Benagerie volcanics, Australia. Phys Earth Planet Inter 173:233–253

    Article  Google Scholar 

  • Worthington PF (2002) A validation criterion to optimize core sampling for the characterization of petrophysical facies. Petrophysics 43:477–493

    Google Scholar 

  • Zelt CA (1999) Modelling strategies and model assessment for wide-angle seismic traveltime data. Geophys J Int 139:183–204

    Article  Google Scholar 

Download references

Acknowledgments

Thanks go to Intrepid Geophysics for technical assistance and access to the 3D Geomodeller API (Application Programming Interface). Special thanks to Philip Chan of the Monash eResearch Centre for his patience in providing technical assistance with Monash Sun Grid computing. Many thanks to Eric de Kemp and his skills with GoCAD. We are grateful for the thorough and constructive reviews from one anonymous reviewer and Li Zhen Cheng, Université du Québec en Abitibi-Témiscamingue. Finally, we thank Guillaume Caumon and Pauline Collon-Drouaillet for their efforts in compiling the Special Publication on 3D Structural Modelling.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Mark Lindsay.

Appendix: Description of Geodiversity Metrics

Appendix: Description of Geodiversity Metrics

The following summary of geodiversity metrics is a compilation of the methods presented in Lindsay et al. (2013a, b).

1.1 A1. Geometrical Geodiversity Metrics

1.1.1 A1.1. Formation Depth and Volume

The shallowest extent of a stratigraphic unit is calculated by determining the depth of the shallowest voxel in the formation under study. The deepest extent is determined from the deepest voxel of a given formation (Fig. 19). Volumes are determined by multiplying the count of stratigraphic unit voxels with the voxel volume, or by simply displaying the voxel count (Fig. 19).

Fig. 19
figure 19

Formation depth and volume calculations. The geological bodies are shown as a set of voxels rather than surfaces, according to the described method. Formation depth analysis detects the location of the deepest or shallowest voxel attributed to a particular geological unit. Volume analysis is performed with a voxel count of the geological unit. Reprinted from Tectonophysics, 594, Lindsay et al. (2013a), Copyright 2013 with permission from Elsevier

1.1.2 A1.2. Average Mean Curvature

Curvature is determined by finding the orientation of a normal section rotated around the surface normal \((\overrightarrow{N})\) at P, a given point on a folded surface (Lisle and Toimil 2007). The maximum magnitude of curvature along the intersection of the folded surface is principal curvature \(k_1\). Curvature of the intersection of the surface and a normal section perpendicular to \(k_1\) is principal curvature \(k_2\) so that \(k_1 > k_2\). Positive principal curvature indicates convex-upward and negative principal curvature indicates concave-upward. Mean curvature (\(M\)) is the arithmetic mean of \(k_1\) and \(k_2\)

$$\begin{aligned} M=\frac{k_1 +k_2 }{2}. \end{aligned}$$
(1)

When \(M >0\) an antiformal (convex) surface is observed, when \(M <0 \) a synformal (concave) surface is observed. \(M = 0\) indicates the surface is a flat plane or a surface where \(k_1 = -k_2\) known as a ‘perfect saddle’ (Lisle and Toimil 2007). Positive Gaussian curvature (\(G)\)

$$\begin{aligned} G=k_1 \cdot k_2 . \end{aligned}$$
(2)

indicates that both principal curvatures \(k_{1}\) and \(k_{2}\) have the same sign and the surface resembles a dome or a basin, if inverted. Negative Gaussian curvature indicates that principal curvatures have different signs and the surface resembles an antiformal or synformal saddle (Fig. 20) (Lisle and Toimil 2007; Mynatt et al. 2007). Please refer to Besl and Jain (1986), Lisle and Robinson (1995) and Lisle and Toimil (2007) for further details and case studies.

Fig. 20
figure 20

Short-distance neighbour calculation constrained to locating the six closest neighbours in a regular grid made of cells with equidistant axes. The sample location assigned with a stratigraphic ID of ‘1’ is surrounded by cells assigned with IDs of ‘1’, ‘2’ and ‘3’ resulting in a neighbourhood relationship metric result of ‘3’. Modified from Lindsay et al. (2013a)

1.1.3 A1.3. Neighbourhood Relationships and Contact Surface Area

Neighbourhood relationships are calculated with a \(k\)-nearest neighbour algorithm (\(k\)-NN). The locations of k nearest neighbours around a point of reference are found, and from that their Euclidean distance can be calculated (Bremner et al. 2005; Friedman et al. 1977). The Euclidean distance is used to constrain which voxels are counted as neighbours and the shortest distances measured along eastings, northings and depth axes resulting in a six-neighbour relationship (Fig. 20). The surface area of the contact between stratigraphic units is identified together with the proportion of overall contact relationships within the three-dimensional volume.

1.2 A2. Geophysical Geodiversity Metrics

1.2.1 A2.1. Root Mean Square

A root-mean-squared value, or ‘rms’, is

$$\begin{aligned} x_\mathrm{rms} =\sqrt{\frac{1}{n}( {x_1^2 +x_2^2 +\cdots +x_n^2 }),} \end{aligned}$$
(3)

where \(x_{1}\), \(x_{2} \ldots x_{n}\) equals the difference between one value and another in a grid.

1.2.2 A2.2. Standard deviation

The standard deviation (\(s\)) of an image is taken as a measure of value spread over a grid

$$\begin{aligned} s=\left( {\frac{1}{n-1}\mathop \sum \limits _{i=1}^n ( {x_i -\bar{x}})^2}\right) ^{\frac{1}{2}}. \end{aligned}$$
(4)

1.2.3 A2.3. Entropy

Information entropy (\(E)\) is used to measure the average bits per pixel over an entire image, representing its global information content (Batty 1974; O’Gorman et al. 2008; Shannon 1948)

$$\begin{aligned} E=\mathop \sum \limits _i^N p_i \log p_i , \end{aligned}$$
(5)

where \(E \) is the sum of all products of \(p\) (probability) of each possible outcome (\(i)\) out of \(N \) total possible outcomes. \(E = 0\) indicates that the image is dominated by large regions of the same value. If in a 1-bit system of two integer values 0 and 1, \(E_\mathrm{max} = 1,\) the image is made of equal proportions of possible values in this case. \(E = 1\) reflects that it is equally likely to find a ‘0’ or a ‘1’ in a given image.

1.2.4 A2.4. Two-Dimensional Correlation Coefficient

Two-dimensional correlation coefficients are typically calculated in geophysical and engineering applications to track changes in two- and three-dimensional objects. The two-dimensional correlation coefficient \(r\) is calculated using

$$\begin{aligned} r=\frac{\mathop \sum \nolimits _m \mathop \sum \nolimits _n ( {A_{mn} -\bar{A} })( {B_{mn} -\bar{B} })}{\sqrt{\mathop \sum \nolimits _m \mathop \sum \nolimits _n ( {( {A_{mn} -\bar{A} })^2})( {\mathop \sum \nolimits _m \mathop \sum \nolimits _n ( {B_{mn} -\bar{B}})^2})} }, \end{aligned}$$
(6)

where \(\bar{A} \) is the global mean of set one and \(\bar{B}\) is the global mean of set two. The purpose of this technique is to identify when patterns in two different sets resemble one another.

1.2.5 A2.3. Hausdorff Distance

The Hausdorff distance measures how far points in two different subsets are from each other. The distance can then be used to understand the level of resemblance two superimposed objects have to each other (Huttenlocher et al. 1993; Olson and Huttenlocher 1997; Rucklidge 1997; Sim et al. 1999; Wang and Suter 2007). The Hausdorff distance, \(d_\mathrm{H}\), between two sets is

$$\begin{aligned} d_\mathrm{H} ( {X,Y})=\max \left\{ {{\begin{array}{l} \mathrm{sup} \\ {x\in X} \\ \end{array} }{\begin{array}{l} \mathrm{inf} \\ {y\in Y} \\ \end{array} }d( {x,y}),\quad {\begin{array}{l} \mathrm{sup} \\ {y\in Y} \\ \end{array} }{\begin{array}{l} \mathrm{inf} \\ {x\in X} \\ \end{array} }d( {x,y})} \right\} , \end{aligned}$$
(7)

where \(X \) and \(Y\) are two non-empty subsets of a metric space (\(M, d)\), ‘sup’ is the supremum, ‘inf ’ is the infimum and \(x \) and \(y\) are points within sets \(X\) and \(Y\), respectively. The supremum defines the upper bounds of subset \(Y\) within set \(X\), whereas the infimum defines the lower bounds of subset \(Y \) within set \(X\). First, determine the shortest distance from point \(x_{1}\) to any point in \(Y,\) then the shortest distance from point \(x_{2}\) to any point in \(Y,\) point \(x_{n}\) to any point in \(Y.\) The Hausdorff distance is the largest distance of those measured from \(x_{1}, x_{2 },{\ldots },x_{n}\).

1.3 A3. Two-Stage Principal Component Analysis Procedure

The following summary of PCA is a compilation of the methods presented in Lindsay et al. (2013a, b). Common sources of model suite variability can be identified by combining metric variability into principal component analyses. Thus, PCA has been chosen to understand the multidimensional problem presented in this manuscript. PCA is a multivariate exploratory data technique that allows complex data interactions to be displayed (Jolliffe 2002). PCA transforms the data orthogonally and highlights relevance by re-organization of the data with respect to the attribute under analysis. Central to PCA is conversion of the potentially correlated original variables, i.e. the geodiversity metrics, into uncorrelated principal components. Data conversion is performed so the first principal component displays the greatest variance, with each subsequent component displaying progressively lower degrees of variance. Each component contains a contribution of variability from all the metrics submitted to the PCA, while each following component contains the next highest possible degree of remaining variance, with the proviso that it is uncorrelated to the preceding components (Jolliffe 2002).

The MATLAB princomp function (http://www.mathworks.com.au/help/toolbox/stats/princomp.html) was used to calculate the coefficients, or loadings, of metric linear combinations. Principal components are calculated in the following manner:

  1. 1.

    Statistical calculations such as the mean, subtraction of deviations from the mean and covariance matrix calculation.

  2. 2.

    Sort the eigenvalues and eigenvectors of the covariance matrix in descending order.

  3. 3.

    Determine eigenvector contribution to eigenvalues.

  4. 4.

    Basis vector determination.

  5. 5.

    Project \(z\)-score-converted original dataset onto the basis vectors.

Loading vectors show the amount of variability any metric has toward a principal component (see Fig. 21). The length of a vector and the angle it has with the principal component axis indicate the strength of the relationship. The longer the vector, the more variability contained within the metric it represents. The smaller the incidence angle of the loading vector to the principal component axis, the more related the vector is to that principal component. Stage one analyses individual stratigraphic units within a metric, to find those which best describe variability within each metric. The first stage is critical to identifying: (1) redundant metrics, and filtering them from analysis and; (2) retaining those metrics that do adequately describe variability in the model suite for analysis in stage two. The filtering of redundant metrics is only performed for those metrics that analyse each stratigraphic unit (depth, volume and short-distance neighbourhood relationship). Metrics that represent the model with a single value, such as the geophysical metrics (refer Sect. A2), are input directly into stage two.

Fig. 21
figure 21

Two-stage PCA method used in this contribution shown as a flowchart. Stage one geodiversity analyses of the volume and depth of formations ‘A’, ‘B’, ‘C’ and ‘D’ are above the solid black line. Note formations are subscripted ‘V’(volume) or ‘D’(depth) to indicate the relevant geodiversity metric. Stage two results are below the black line and are where all the metrics (volume and depth, in this example) are simultaneously examined with PCA. Note that any metrics, not just depth and volume, can be included in stage two. The volume of formations B and C (circled) and the depth of formations A and D are shown in stage one to contribute most to model suite variability. The most influential metric for model suite variability, the depth of formation A, is revealed in stage two. Reprinted from Tectonophysics, 594, Lindsay et al. (2013a), Copyright 2013 with permission from Elsevier

Stage two repeats the PCA process, with all the metrics described in Sects. A2 and A3. A simplified example of the two-stage PCA process is shown in Fig. 21. Stratigraphic units B and C contribute most to the first principal component and second component variability respectively for the volume metric. Units A and D contribute most variability to the first and second principal components respectively for the depth geodiversity metric. Units B and C represent the volume metrics, and units A and D represent the depth metrics. Figure 21 shows that in stage two, where the volume and depth metrics are combined, depth of unit A has been determined to be the most influential in terms of model suite variability.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lindsay, M., Perrouty, S., Jessell, M. et al. Inversion and Geodiversity: Searching Model Space for the Answers. Math Geosci 46, 971–1010 (2014). https://doi.org/10.1007/s11004-014-9538-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11004-014-9538-x

Keywords

Navigation