Hostname: page-component-848d4c4894-8bljj Total loading time: 0 Render date: 2024-06-17T08:41:01.508Z Has data issue: false hasContentIssue false

Extraction of 40Ar-39Ar Ages from a Multicomponent Mixture: A Case Study From the Tatra Mountains, Poland

Published online by Cambridge University Press:  01 January 2024

Artur Kuligiewicz*
Affiliation:
Institute of Geological Sciences, Polish Academy of Sciences, Senacka 1, 31-002 Kraków, Poland
Michał Skiba
Affiliation:
Institute of Geological Sciences, Jagiellonian University, Gronostajowa 3a, 30-387 Kraków, Poland
Marek Szczerba
Affiliation:
Institute of Geological Sciences, Polish Academy of Sciences, Senacka 1, 31-002 Kraków, Poland
Chris M. Hall
Affiliation:
Department of Earth and Environmental Sciences, University of Michigan, Ann Arbor, MI 48109-1005, USA
Dorota Bakowska
Affiliation:
Institute of Geological Sciences, Polish Academy of Sciences, Senacka 1, 31-002 Kraków, Poland

Abstract

Extraction of meaningful information on the timing of fault activity from clay gouges using radiometric dating methods, such as those based on the K-Ar system, can be challenging. One of the factors complicating interpretation of the radiometric dating results is the presence of multiple K-bearing components in the gouge material. In the current study, an attempt was made to develop a new interpretative method for K-Ar and 40Ar-39Ar dating, capable of handling a three-component mixture. In addition, the mineral composition of clay gouges from the Tatra Mountains (Poland), which has not been investigated before, is reported. The mineral compositions of the bulk clay gouge material and separated size fractions were determined by X-ray diffractometry and Fourier-transform infrared spectroscopy. The gouge samples were composed of quartz, dioctahedral mica (as a discrete phase and as a component of mixed-layered illite-smectite), and chlorite, commonly with plagioclase and more rarely with K-feldspar, dioctahedral smectite, calcite, anatase, or trace kaolinite. One feldspar-free sample containing three mica polytypes (1Md, 1M, and 2M1) was chosen for dating with the 40Ar-39Ar method. The results of 40Ar-39Ar dating were interpreted using three concepts: Illite Age Analysis (IAA), a method based on the MODELAGE software, and a newly developed three-component concept. The age values obtained with IAA were −14 Ma ± 31 Ma and 180 ± 91 Ma for authigenic (1Md) and inherited (1M + 2M1) components, respectively. The MODELAGE-based approach returned –4 ± 40 Ma and 165 ± 62 Ma. The three-component approach returned age values of polytypes as follows: 1Md, 15 ± 37 Ma; 1M, 135 ± 57 Ma; 2M1, 121 ± 56 Ma based on the medians and the interquartile ranges of non-normal distributions of Monte Carlo-simulated age values. The results obtained indicated that the 1Md polytype was probably formed during the most recent stage of fault activity, while 1M and 2M1 polytypes are of equal age, roughly.

Type
Original Paper
Copyright
Copyright © The Author(s), under exclusive licence to The Clay Minerals Society 2022

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

Footnotes

This article was updated to correct errors in the formatting of Equations 5, 6, 7, and 8 that were introduced during the production process.

References

Abad, I., Jiménez-Millán, J., Schleicher, A. M., & van der Pluijm, B. A. (2017). Mineral characterization, clay quantification and Ar-Ar dating of faulted schists in the Carboneras and Palomares Faults (Betic Cordillera, SE Spain). European Journal of Mineralogy, 29(1), 1734.CrossRefGoogle Scholar
Anczkiewicz, A. A., Danišik, M., & Środoń, J. (2015). Multiple low-temperature thermochronology constraints on exhumation of the Tatra Mountains: New implication for the complex evolution of the Western Carpathians in the Cenozoic. Tectonics, 34(11), 22962317.CrossRefGoogle Scholar
Bense, F. A., Wemmer, K., Löbens, S., & Siegesmund, S. (2014). Fault gouge analyses: K–Ar illite dating, clay mineralogy and tectonic significance—a study from the Sierras Pampeanas, Argentina. International Journal of Earth Sciences, 103(1), 189218.CrossRefGoogle Scholar
Boles, A., Mulch, A., & van der Pluijm, B. (2018). Near-surface clay authigenesis in exhumed fault rock of the Alpine Fault Zone (New Zealand); O-H-Ar isotopic, XRD and chemical analysis of illite and chlorite. Journal of Structural Geology, 111, 2741.CrossRefGoogle Scholar
Brodie, K., Fettes, D., Harte, B., & Schmid, R. (2004). A systematic nomenclature for metamorphic rocks. 3. Structural terms including fault-rock terms. Recommendations by the IUGS Subcommission on the systematics of metamorphic rocks. SCMR website (www. bgs. ac. uk/SCMR).Google Scholar
Burda, J., Gawęda, A., & Klötzli, U. (2011). Magma hybridization in the Western Tatra Mts. granitoid intrusion (S-Poland, Western Carpathians). Mineralogy and Petrology, 103, 1936.CrossRefGoogle ScholarPubMed
Burda, J., Gawęda, A., & Klötzli, U. (2013). U-Pb zircon age of the youngest magmatic activity in the High Tatra granites (Central Western Carpathians). Geochronometria, 40(2), 134144.CrossRefGoogle Scholar
Clauer, N. (2013). The K-Ar and 40Ar/39Ar methods revisited for dating fine-grained K-bearing clay minerals. Chemical Geology, 354, 163185.CrossRefGoogle Scholar
Clauer, N., Zwingmann, H., Liewig, N., & Wendling, R. (2012). Comparative Ar/Ar and K-Ar dating of illite-type clay minerals: A tentative explanation for age identities and differences. Earth-Science Reviews, 115(1–2), 7696.CrossRefGoogle Scholar
Collins, G. (1990). Fundamental numerical methods and data analysis. Published by the author.Google Scholar
Doebelin, N., & Kleeberg, R. (2015). Profex: a graphical user interface for the Rietveld refinement program BGMN. Journal of Applied Crystallography, 48, 15731580.CrossRefGoogle ScholarPubMed
Dong, H., Hall, C. M., Peacor, D. R., & Halliday, A. N. (1995). Mechanisms of argon retention in clays revealed by laser Ar- Ar dating. Science, 267(5196), 355359.CrossRefGoogle Scholar
Dong, H., Hall, C. M., Halliday, A. N., Peacor, D. R., Merriman, R. J., & Roberts, B. (1997). 40Ar/39Ar illite dating of Late Caledonian (Acadian) metamorphism and cooling of K-bentonites and slates from the Welsh Basin, U.K. Earth and Planetary Science Letters, 150(3-4), 337351.CrossRefGoogle Scholar
Dong, H., Hall, C. M., Peacor, D. R., Halliday, A. N., & Pevear, D. R. (2000). Thermal Ar/Ar separation of diagenetic from detrital illitic clays in Gulf Coast shales. Earth and Planetary Science Letters, 175(3–4), 309325.CrossRefGoogle Scholar
Drits, V. A., & Tchoubar, C. (1990). X-ray diffraction by disordered lamellar structures. Springer.CrossRefGoogle Scholar
Eberl, D. D., Środoń, J., Lee, M., Nadeau, P. H., & Northorp, H. R. (1987). Sericite from the Silverton Caldera, Colorado: Correlation among structure, composition, origin, and particle thickness. American Mineralogist, 72, 914934.Google Scholar
Fitz-Díaz, E., Hall, C. M., & van der Pluijm, B. A. (2016). XRD-based 40Ar/39Ar age correction for fine-grained illite, with application to folded carbonates in the Monterrey Salient (northern Mexico). Geochimica et Cosmochimica Acta, 181, 201216.CrossRefGoogle Scholar
Gawęda, A. (2007). Variscan granitoid magmatism in Tatra Mountains–the history of subduction and continental collision. Granitoids in Poland, AM Monograph No., 1, 319332.Google Scholar
Gawęda, A., & Włodyka, R. (2012). The origin of post-magmatic Ca-Al minerals in granite-diorite mingling zones: The Tatra granitoid intrusion, Western Carpathians. Neues Jahrbuch für Mineralogie-Abhandlungen, 190, 2947.CrossRefGoogle Scholar
Gawęda, A., Doniecki, T., Burda, J., & Kohút, M. (2005). The petrogenesis of quartz-diorites from the Tatra Mountains (Central Western Carpathians): an example of magma hybridisation. Neues Jahrbuch Fur Mineralogie-Abhandlungen, 181(1), 95109.CrossRefGoogle Scholar
Gawęda, A., Szopa, K., & Chew, D. (2014). LA-ICP-MS U-Pb dating and REE patterns of apatite from the Tatra Mountains, Poland as a monitor of the regional tectonomagmatic activity. Geochronometria, 41(4), 304314.CrossRefGoogle Scholar
Goddard, J. V., & Evans, J. P. (1995). Chemical-changes and fluid-rock interaction in faults of crystalline thrust sheets, northwestern Wyoming, USA. Journal of Structural Geology, 17(4), 533547.CrossRefGoogle Scholar
Grathoff, G. H., & Moore, D. M. (1996). Illite polytype quantification using WILDFIRE© calculated X-ray diffraction patterns. Clays and Clay Minerals, 44(6), 835842.CrossRefGoogle Scholar
Haines, S. H., & van der Pluijm, B. A. (2008). Clay quantification and Ar-Ar dating of synthetic and natural gouge: Application to the Miocene Sierra Mazatan detachment fault, Sonora, Mexico. Journal of Structural Geology, 30(4), 525538.CrossRefGoogle Scholar
Haines, S. H., & van der Pluijm, B. A. (2012). Patterns of mineral transformations in clay gouge, with examples from low-angle normal fault rocks in the western USA. Journal of Structural Geology, 43, 232.CrossRefGoogle Scholar
Hall, C. M. (2014). Direct measurement of recoil effects on 40Ar/39Ar standards. In Jourdan, F., Mark, D. F., & Verati, C. (eds), Advances in Ar/Ar Dating: From Archaeology to Planetary Sciences (pp. 5362). Geological Society Special Publications 378(1), Geological Society, London.Google Scholar
Hayman, N. W. (2006). Shallow crustal fault rocks from the Black Mountain detachments, Death Valley, CA. Journal of Structural Geology, 28(10), 17671784.CrossRefGoogle Scholar
Jackson, M. L. (1969). Soil Chemical Analysis-Advanced Course. Madison, WI: published by the author.Google Scholar
Jurewicz, E. (2005). Geodynamic evolution of the Tatra Mts. and the Pieniny Klippen Belt (Western Carpathians): problems and comments. Acta Geologica Polonica, 55(3), 295338.Google Scholar
Jurewicz, E. (2006). Petrophysical control on the mode of shaering in the sedimentary rocks and granitoid core of the Tatra Mountains during Late Cretaceous nappe-thrusting and folding, Carpathians, Poland. Acta Geologica Polonica, 56(2), 159170.Google Scholar
Jurewicz, E., & Bagihski, B. (2005). Deformation phases in the selected shear zones within the Tatra Mountains granitoid core. Geologica Carpathica, 56, 1728.Google Scholar
Jurewicz, E., & Kozłowski, A. (2003). Formation conditions of quartz mineralization in the mylonitic zones and on the slickenside fault planes in the High Tatra granitoids. Archiwum Mineralogiczne, 54, 6576.Google Scholar
Kelley, S. (2002). K-Ar and Ar-Ar dating. Reviews in Mineralogy and Geochemistry, 47(1), 785818.CrossRefGoogle Scholar
Kohút, M., & Sherlock, S. C. (2003). Laser microprobe 40Ar-39Ar analysis of pseudotachylyte and host-rocks from the Tatra Mountains, Slovakia: evidence for late Palaeogene seismic/tectonic activity. Terra Nova, 15(6), 417424.CrossRefGoogle Scholar
Králiková, S., Vojtko, R., Sliva, U., Minár, J., Fügenschuh, B., Kovác, M., & Hók, J. (2014). Cretaceous—Quaternary tectonic evolution of the Tatra Mts (Western Carpathians): constraints from structural, sedimentary, geomorphological, and fission track data. Geologica Carpathica, 65(4), 307326.CrossRefGoogle Scholar
Kuligiewicz, A., Derkowski, A., Szczerba, M., Gionis, V., & Chryssikos, G. D. (2015). Revisiting the infrared spectrum of the water-smectite interface. Clays and Clay Minerals, 63(1), 1529.CrossRefGoogle Scholar
Leichmann, J., Jacher-Sliwczynska, K., & Broska, I. (2009). Element mobility and fluid path ways during feldspar alteration: textural evidence from cathodoluminescece and electron microprobe study of an example from tonalites (High Tatra, Poland-Slowakia). Neues Jahrbuch für Mineralogie –Abhandlungen, 186, 110.CrossRefGoogle Scholar
Maluski, H., Rajlich, P., & Matte, P. (1993). 40Ar-39Ar dating of the Inner Carpathians Variscan basement and Alpine mylonitic overprinting. Tectonophysics, 223(3-4), 313337.CrossRefGoogle Scholar
Mancktelow, N., Zwingmann, H., & Mulch, A. (2016). Timing and conditions of clay fault gouge formation on the Naxos detachment (Cyclades, Greece). Tectonics, 35(10), 23342344.CrossRefGoogle Scholar
McCarty, D. K., Sakharov, B. A., & Drits, V. A. (2009). New insights into smectite illitization: A zoned K-bentonite revisited. American Mineralogist, 94(11-12), 16531671.CrossRefGoogle Scholar
Moore, D., & Reynolds, R. (1997). X-ray Diffraction and the Identification and Analysis of Clay Minerals. Oxford University Press.Google Scholar
Peacor, D. R., Bauluz, B., Dong, H., Tillick, D., & Yonghong, Y. (2002). Transmission and analytical electron microscopy evidence for high Mg contents of 1M illite: absence of 1M polytypism in normal prograde diagenetic sequences of pelitic rocks. Clays and Clay Minerals, 50(6), 757765.CrossRefGoogle Scholar
Pevear, D. R. (1992). Illite age analysis, a new tool for basin thermal history analysis. In Kharaka, Y. K. & Maest, A. S. (Eds.), Water-rock interaction (pp. 12511254). AA Balkema.Google Scholar
Pevear, D. R. (1999). Illite and hydrocarbon exploration. Proceedings of the National Academy of Sciences, 96, 34403446.CrossRefGoogle ScholarPubMed
Piotrowska, K. (1970). Fotointerpretacja i geneza struktur nieciagłych w polskiej cze_sci Tatr Wysokich. Acta Geologica Polonica, 20, 365411 [in Polish].Google Scholar
Plašienka, D., Grecula, P., Putiš, M., Kovác, M., & Hovorka, D. (1997). Evolution and structure of the Western Carpathians: an overview. In Grecula, P., Hovorka, D., & Putiš, M. (Eds.), Geological evolution of the Western Carpathians (pp. 124). Mineralia Slovaca – Monograph.Google Scholar
Russell, J. D., & Farmer, V. C. (1964). Infra-red spectroscopic study of the dehydration of montmorillonite and saponite. Clay Minerals Bulletin, 5(32), 443464.CrossRefGoogle Scholar
Samson, S. D., & Alexander, E. C. (1987). Calibration of the interlaboratory 40Ar- 39Ar dating standard, MMhb-1. Chemical Geology: Isotope Geoscience section, 66(1), 2734.Google Scholar
Schleicher, A. M., Warr, L. N., & van der Pluijm, B. (2009). On the origin of mixed-layered clay minerals from the San Andreas Fault at 2.5-3 km vertical depth (SAFOD drillhole at Parkfield, California). Contributions to Mineralogy and Petrology, 157(2), 173187.CrossRefGoogle Scholar
Sibson, R. (1977). Fault rocks and fault mechanisms. Journal of the Geological Society, 133(3), 191213.CrossRefGoogle Scholar
Skiba, M. (2007). Clay mineral formation during podzolization in an alpine environment of the Tatra Mountains, Poland. Clays and Clay Minerals, 55(6), 618634.CrossRefGoogle Scholar
Śmigielski, M., Sinclair, H. D., Stuart, F. M., Persano, C., & Krzywiec, P. (2016). Exhumation history of the Tatry Mountains, Western Carpathians, constrained by low-temperature thermochronology. Tectonics, 35(1), 187207.CrossRefGoogle Scholar
Solum, J. G., van der Pluijm, B. A., & Peacor, D. R. (2005). Neocrystallization, fabrics and age of clay minerals from an exposure of the Moab Fault, Utah. Journal of Structural Geology, 27(9), 15631576.CrossRefGoogle Scholar
Środoń, J. (1999). Use of clay minerals in reconstructing geological processes: recent advances and some perspectives. Clay Minerals, 34(1), 2737.CrossRefGoogle Scholar
Środoń, J. (2013). Identification and Quantitative Analysis of Clay Minerals. In Bergaya, F., & Lagaly, G. (eds), Handbook of Clay Science (pp. 2549). Developments in Clay Science 5, Elsevier, Amsterdam.CrossRefGoogle Scholar
Środoń, J., Zeelmaekers, E., & Derkowski, A. (2009). The charge of component layers of illite-smectite in bentonites and the nature of end-member illite. Clays and Clay Minerals, 57(5), 649671.CrossRefGoogle Scholar
Steiger, R. H., & Jäger, E. (1977). Subcommission on geochronology: Convention on use of decay constants in geo- and cosmochronology. Earth and Planetary Science Letters, 36(3), 359362.CrossRefGoogle Scholar
Surace, I. R., Clauer, N., Thélin, P., & Pfeifer, H.-R. (2011). Structural analysis, clay mineralogy and K-Ar dating of fault gouges from Centovalli Line (Central Alps) for reconstruction of their recent activity. Tectonophysics, 510(1-2), 8093.CrossRefGoogle Scholar
Szczerba, M., & Środoń, J. (2009). Extraction of diagenetic and detrital ages and of the 40Kdetrital/40 Kdiagenetic ratio from K-Ar dates of clay fractions. Clays and Clay Minerals, 57(1), 93103.CrossRefGoogle Scholar
Torgersen, E., Viola, G., Zwingmann, H., & Harris, C. (2014). Structural and temporal evolution of a reactivated brittleductile fault - Part II: Timing of fault initiation and reactivation by K-Ar dating of synkinematic illite/muscovite. Earth and Planetary Science Letters, 407, 221233.CrossRefGoogle Scholar
Turnau-Morawska, M. (1948). Z mikrogeologii trzonu krystalicznego Tatr (Microgeological researches in central part of the crystaline Tatra). Kosmos seriaA, 65, 59100. In Polish with English summary.Google Scholar
Uchman, A. & Michalik, M. (1997) Podłoże geologiczne In: Plan Ochrony TPN. Operat ochrona przyrody nieożywionej i gleb. S. Skiba, A. Kotarba (Ed.). Zakopane: Tatrzański Park Narodowy [in Polish]Google Scholar
Ufer, K., Kleeberg, R., Bergmann, J., & Dohrmann, R. (2012). Rietveld refinement of disordered illite-smectite mixed-layer structures by a recursive algorithm. I: One-dimensional patterns. Clays and Clay Minerals, 60(5), 507534.CrossRefGoogle Scholar
van der Pluijm, B., & Hall, C. M. (2015). Fault Zone (Thermochronology). In Rink, W. Jack & Thompson, J. W. (Eds.), Encyclopedia of Scientific Dating Methods (pp. 269274). Springer Netherlands.CrossRefGoogle Scholar
van der Pluijm, B. A., Hall, C. M., Vrolijk, P. J., Pevear, D. R., & Covey, M. C. (2001). The dating of shallow faults in the Earth's crust. Nature, 412(6843), 172175.CrossRefGoogle ScholarPubMed
Viola, G., Zwingmann, H., Mattila, J., & Kapyaho, A. (2013). K-Ar illite age constraints on the Proterozoic formation and reactivation history of a brittle fault in Fennoscandia. Terra Nova, 25(3), 236244.CrossRefGoogle Scholar
Yan, Y., Tillick, D., Peacor, D. R., & Simmons, S. F. (2001). Genesis of dioctahedral phyllosilicates during hydrothermal alteration of volcanic rocks: II The Broadlands-Ohaaki hy-drothermal system, New Zealand. Clays and Clay Minerals, 49(2), 141155.CrossRefGoogle Scholar
Ylagan, R. F., Pevear, D. R., & Vrolijk, P. J. (2000). Discussion of “Extracting K-Ar ages from shales: a theoretical test”. Clay Minerals, 35(3), 599604.CrossRefGoogle Scholar
Ylagan, R. F., Kim, C. S., Pevear, D. R., & Vrolijk, P. J. (2002). Illite polytype quantification for accurate K-Ar age determination. American Mineralogist, 87(11-12), 15361545.CrossRefGoogle Scholar
York, D., Evensen, N. M., Lopez Martinez, M., & De Basabe Delgado, J. (2004). Unified equations for the slope, intercept, and standard errors of the best straight line. American Journal of Physics, 72(3), 367375.CrossRefGoogle Scholar
Supplementary material: File

Kuligiewicz et al. supplementary material
Download undefined(File)
File 10.4 KB
Supplementary material: File

Kuligiewicz et al. supplementary material
Download undefined(File)
File 100.8 KB
Supplementary material: File

Kuligiewicz et al. supplementary material
Download undefined(File)
File 185.7 KB
Supplementary material: File

Kuligiewicz et al. supplementary material
Download undefined(File)
File 295.1 KB