Next Article in Journal
Adsorption of Polyions on Flat TiO2 Surface
Next Article in Special Issue
Large Amount of Excess Argon in Hydrothermal Quartz from the Vangtat Orogenic Gold Belt, Southern Laos: New In-Sight from K-Ar and Noble Gas Isotope Analyses
Previous Article in Journal
Textures and Chemical Compositions of Nb-Bearing Minerals and Nb Mineralization in the Shuangshan Nepheline Syenite Pluton, East Qinling, China
Previous Article in Special Issue
Formation of a Composite Albian–Eocene Orogenic Wedge in the Inner Western Carpathians: P–T Estimates and 40Ar/39Ar Geochronology from Structural Units
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Illite-Age-Analysis (IAA) for the Dating of Shallow Faults: Prerequisites and Procedures for Improvement

Department of Earth System Sciences, Yonsei University, 50 Yonsei-ro, Seoul 120-749, Korea
*
Author to whom correspondence should be addressed.
Minerals 2021, 11(11), 1162; https://doi.org/10.3390/min11111162
Submission received: 30 August 2021 / Revised: 18 October 2021 / Accepted: 19 October 2021 / Published: 21 October 2021
(This article belongs to the Special Issue Frontier of the K–Ar (40Ar/39Ar) Geochronology)

Abstract

:
Fault age determination using the illite-age-analysis (IAA) method for fault gouges has played a key role in providing absolute age information in tectonic evolution studies for the last 20 years. The accuracy and precision of the IAA method depend on (1) how to reasonably quantify the relative content of 1M/1Md illite generated from fault activity compared to detrital 2M1 illite in the size fractions of the fault gouge, and (2) how to minimize the error factors in K-Ar or Ar-Ar dating analysis. XRD-based quantitative analysis of illite polytype has made great progress in accuracy by generating a simulated XRD pattern of 1M/1Md polytype using WILDFIRE© and full-pattern-fitting it with the XRD pattern measured from size fractions of the fault gauge. Nevertheless, the results of quantitative analysis of illite polytypes may vary depending on the sample state of the size fractions for XRD analysis, especially the preferred orientation due to the layered crystal structure of illite. In addition, the radiometric dating results may be distorted depending on the error factor of the dating method itself and on the mineral composition of the size fractions, that is, the presence of K-containing minerals such as biotite and K-feldspar other than illite. In this study, we reviewed previous studies that determined fault activity ages by applying IAA to fault gouges. From this, the prerequisites and recommendations for each of the five steps (particle size separation process, XRD analysis process, polytype quantification, radiometric dating, IAA plot) for improving the IAA method are summarized and presented. The continuous application of the improved IAA is expected to greatly contribute to the study of tectonic evolution through geological time.

1. Introduction

The Illite-Age-Analysis (IAA) method was first proposed by Pevear (1992, 1999) [1,2] for the catalytic dating of sedimentary basins. After van der Pluijm et al. (2001) discovered a new application for defining the age of fault-thrust development with the IAA method [3], it has been applied to the shallow faults of various tectonic environments by a number of researchers for the past 20 years, and has played a decisive role in the study of tectonic evolution and understanding of seismic phenomena. In particular, the development of the WILDFIRE© program by Reynolds (1994) [4] has made great strides in the quantitative analysis of illite polytype by simulating the 1M/1Md polytype patterns. Moreover, the incorporation of micro-encapsulation in the 40Ar-39Ar method [3] and the improvement of the K-Ar method with a small amount of sample have greatly enhanced the reliability of the IAA method.
The relative content of illite polytype is a key variable that determines the reliability of the IAA method. Most researchers have applied X-ray diffraction (XRD) analysis and the illite quantification method based on the WILDFIRE© program, but there are some differences in the XRD analysis conditions and the method of using the simulated pattern generated by WILDFIRE©. For example, since the XRD pattern of illite has a layer structure, the relative intensity of peaks may be distorted due to the preferred orientation of the particles. However, the back-/side-packing method applied in most studies to minimize this effect is difficult to consistently guarantee the state of the analyzed sample for each researcher, so it may become an error factor in the quantitative analysis value. This will be discussed further in Section 4. As another example, in the WILDFIRE©-based quantitative analysis method used in most IAA studies, there is a difference depending on the researcher in the method of using the simulated XRD pattern. This will be discussed further in Section 5.
Furthermore, there may be some error factors in the process of determining the absolute age of each particle size. Both the radiometric K-Ar or Ar-Ar methods have advantages and disadvantages, and this problem is still debated. This will be discussed further in Section 6. In addition, the state of each fraction, such as the presence of K-containing minerals other than illite, and the presence of K+ in the exchangeable site of layer silicates, can also be an important factor that can affect the dating results.
Although there are fault activity dating methods such as U-Pb dating and Rb-Sr dating for carbonate minerals, the IAA method, which has a wider application range, is still a highly useful method for determining the absolute age of a shallow fault activity. Therefore, in order to obtain reasonable and consistent dating results by applying it, a systematic process should be established in consideration of the aforementioned influencing factors (i.e., mineralogy of sample, error factors in the particle size separation process, preferred orientation of the sample, polytype quantification method, selection of dating method, etc.). In this paper, we summarize the studies on fault dating using the IAA method, and review the differences in the applied quantification method in detail and their results. Based on this, we would like to suggest a direction for improvement and a systematic IAA application procedure to minimize the error and error range of the IAA chronology and to obtain reliable results.

2. Basic Concept of IAA and Previous Studies

Fault gouges, a product of fault activity, generally appear as a mixture of 1M/1Md illite generated due to fault activity and detrital 2M1 illite derived from surrounding rocks [3]. This is a factor that makes it difficult to determine the age of fault activity using fault gouges. IAA is a method proposed to solve this problem, and to determine the generation age of only 1M/1Md illite generated by fault activity. The basic concept of IAA is to obtain the y-axis intercept of 0% detrital 2M1 illite from linear extrapolation through a graphical plot of the dating data (y-axis) of three or more size fractions separated from one fault gouge, versus the relative content of 2M1 illite of each fraction (x-axis) in the binary system between 1M/1Md − 2M1 illite.
When the K-Ar dating formula is arranged, the linear relationship formula is obtained between the term of the dating (eλt − 1) and the mixing rate (x) of detrital mica as seen in the following equation. This is only possible under conditions where K-concentrations of 1M/1Md and 2M1 illite are equal.
−eλt − 1 = (λe + λβ)Ar/λe K
       = (λe + λβ)(Ara(1 − x) + Ard·x)/λe·K
       = (λe + λβ)(Ara + (Ard − Ara)x)/λe·K
Ar(a): Radiogenic argon content of mica clay minerals formed by fault activity.
Ar(d): Radiogenic argon content of detrital mica Argon content.
The y-axis intercept is determined as the generation age of 1M/1Md illite, that is, a fault activity age. A conceptual diagram of the IAA method is presented in Figure 1.
Therefore, the accuracy and precision of the IAA method depend on (1) how to reasonably quantify the relative content of 1M/1Md illite generated from fault activity compared to detrital 2M1 illite in the size fractions of the fault gouge, and (2) how to minimize the error factors in K-Ar or Ar-Ar radiometric dating analysis. In the 20 years since it was first applied to shallow faults by van der Pluijm et al. (2001) [3], many fault-dating studies have been conducted. Table 1 lists previous studies using IAA and the respective experimental and methodological setup, including selected size fractions, XRD conditions (type of equipment, aluminum holder/capillary tube, detector type, etc.), illite polytype quantification method, and dating method for each study result.
In most studies, <2 μm particle size was separated into 3 to 4 particle size fractions [3,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27], but in some studies, >2 μm fraction was also separated [28,29,30,31,32]. The particle size range for each fraction is slightly different depending on the research (Table 1). The XRD equipment used in most studies is the conventional powder diffractometry, and it seems to have been loaded by back/side-packing the powder sample in an aluminum holder and measured [3,5,6,7,8,9,10,11,12,17,18,21,25,27,28,29,31]. Contrary to this, some studies used capillary tubes as sample holders to minimize the preferred orientation effect of grains [13,14,15,16,19,20,22,23,24,26,30,32]. Illite polytype quantification is the most important factor in determining the reliability of IAA results, but there are differences among researchers in the experimental set-ups of quantitative analysis. Therefore, each experimental set-up applied in the IAA process will be discussed in more detail below. Several methods have been proposed so far, and most are based on simulated XRD patterns generated with WILDFIRE© [3,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,25,26,27,30,31,32]. Both K-Ar and Ar-Ar methods were used as radiometric dating methods (Table 1).

3. Selection of Size Fractions and Its Interval

The key to particle size separation for IAA application is to make the interval with difference in the relative content of illite polytypes for each particle size fraction, and to separate finer fractions with a higher 1M/1Md illite content as much as possible. In some studies, even the extremely fine fraction of <0.02 or <0.05 was separated [3,5,6,8,9,10,11,12,17,18,24,25]. In this case, it contains almost pure 1M/1Md illite, so it may show a value close to the fault activity age, and the error can be reduced when applying IAA. In order to secure the amount of sample required for polytype quantification and dating of the extremely fine fraction, it takes a lot of time for particle size separation using a high-speed centrifuge (SIGMA 4-16S, SIGMA, Darmstadt, Germany).
If micro-focused XRD and thin capillary tubes are used in XRD analysis, only a few mg of sample is required, and the time required for particle size separation can be greatly saved because only the amount of sample required for radiometric dating is secured. In addition, in previous studies, it has been reported that the mineral composition of <0.1 μm particle size is more than 90% 1M/1Md illite [13,14,15,16,19,23,26]. The time required for separation of the <0.1 μm particle size fraction using the centrifuge method is much shorter than that of the <0.02 and <0.05 μm particle size fractions. Therefore, considering the amount of sample required for XRD analysis, the time required for particle size separation, and the 1M/1Md illite content, it seems appropriate to set the smallest particle size fraction to <0.1 μm.
In some studies, a particle size range of >2 μm is analyzed [28,29,30,31,32], but a particle size of >2 μm contains a large amount of other minerals, which becomes an error factor in quantitative analysis of illite polytype. In particular, at a particle size of >2 μm, K-containing minerals such as K-feldspar and biotite are included in some cases [28,29,30,31,32], which may affect the dating value, thereby reducing the reliability of IAA. Therefore, it would be appropriate to select three or more fractions at a particle size of <2 μm for IAA.

4. X-ray Diffractometry Procedure for IAA

The key to XRD analysis of size fractions is how to obtain all (hkl) reflections of the illite polytype with an ideal peak-intensity ratio. This indicates that XRD analysis should be performed by minimizing the inevitable preferred orientation effect in the layer crystal structure. Therefore, samples for XRD analysis should be loaded as randomly as possible.
The back-/side-packing method using the aluminum holder increases randomness rather than the simple top-packing method. However, because of the difference in the amount of sample per unit volume used and the packing strength for each case, there is an inevitable difference in the degree of randomness for each case.
On the other hand, the capillary tube can maximize randomness and reduce the difference between researchers. In addition, the capillary tube can be analyzed with only a small amount of sample, so it is more useful for fine-size fractions where it is difficult to secure a sufficient amount of sample.
Although the type of XRD equipment is less important, micro-focused XRD equipment with a 2D-detector (image plate) can obtain an XRD pattern with good peak-selectivity, targeting a microscopic area of a thin capillary tube even with an extremely small amount of sample. Therefore, the micro-focused XRD equipment is optimized for XRD analysis for IAA, and the accuracy and precision of illite polytype quantification results can be enhanced. Indeed, Song et al. (2014) [14] successfully obtained high-resolution (hkl) reflections in a random state for the first time using micro-focused XRD equipment with a 2D-image plate attached to an extremely small amount of sample loaded into a thin capillary tube (0.6 mm in diameter). This method has been applied recently in several studies [13,14,15,16,20,22,23,26].

5. WILDFIRE©-Based Polytype Quantification

5.1. Simulation of Polytype XRD Patterns Using WILDFIRE©

The randomly-mounted measured XRD pattern is a mixture of reflections of 1M/1Md and 2M1 illite polytypes. For this pattern, the relative content of each polytype must be determined before IAA can be applied. Illite, a layer silicate, has many factors that affect the relative peak intensity of (hkl) reflections, such as crystallinity, stacking ordering of layers, and interlayer expandability, etc., as well as the preferred orientation due to the layer structure [4]. For this reason, it is difficult to accurately determine the relative content of illite polytypes by applying a general XRD-based quantitative analysis method.
To overcome this problem, in most previous studies, polytype simulated XRD patterns were created using WILDFIRE© developed by Reynolds (1994) [4] and used for the quantitative analysis of clay minerals. WILDFIRE©, a forward model algorithm, can create various types of 3-dimensional simulated patterns by using crystallographic parameters affecting the XRD pattern of illite polytype as variables. In the WILDFIRE©-based quantitative analysis method, an appropriate pattern is selected and used through iteration that repeats the process of creating a pattern with different variables.
WILDFIRE© is very useful for creating simulated patterns of 1Md polytypes, especially with low crystallinity and poor regularity in the stacking type of layer structure. Since the simulated patterns of 1M and 2M1 consider only a few parameters, such as crystallinity and trans/cis octahedral sheet, it is not difficult to determine a representative simulated pattern. On the other hand, in the case of 1Md polytype, there are many crystallographic parameters that affect the simulated pattern. WILDFIRE© is designed to reflect these parameters and create simulated patterns for various combinations of each parameter variable. The parameters considered as variables to create a simulated pattern of 1Md polytype in WILDFIRE© are as follows.
-
probability of zero rotation (P0)
-
probability of 120 rotation (P120)
-
fraction of n.60 degree rotation (F60)
-
proportion of cis-vacant layers (Pcis)
-
mean defect-free (Coherence) distance (MDFD)
-
water in expandable interlayers
-
crystallite thickness (% expandability)
-
no. of unit cells along X (N1)
-
no. of unit cells along Y (N2)
-
no. of unit cells along Z (N3)
-
K and Fe fraction in the structure
-
Randomness of sample (Dollase factor)
-
ordering of the illite/smectite (Reichweite), etc.
2M1 and 1Md illite simulated patterns under various conditions created by WILDFIRE© using the above parameters as variables provide core basic data for the developed illite polytype quantification method [8,19,33,34], etc. Boles et al. (2018) [35] suggested a WILDFIRE© Model End-member library by creating 20 patterns for 2M1 illite and 695 patterns for 1Md illite using these parameters as variables, respectively.

5.2. Illite Polytype Quantification

For Illite polytype quantification, the previously introduced WILDFIRE©-based quantification method is most commonly used. In addition, there are polytype end-member standards methods [24,31] and methods based on Rietveld refinement [28].
Two main types of quantitative analysis of illite polytype based on WILDFIRE© has been developed as follows; (1) A method using the area ratio of polytype-specific peaks in simulated patterns of 2M1 and 1M/1Md polytypes made by WILDFIRE© modeling [33], and (2) quantification method through graphically best-fitting ratio between mixed pattern made with simulated patterns of illite polytypes and measured pattern [14,33,34]. The first method proposed by Grathoff and Moore (1996) [33] is that in the simulated patterns created with WILDFIRE©, the relative area ratio is calculated for each of the five unique peaks of 2M1 illite against the area of the 2.58 Å/35° 2θ (Cu Kα) peak, which is the common peak of 2M1 and 1Md illite. A linear equation between the 2M1 content and the area ratios is then derived, and then the 2M1 content in a natural sample is determined by substituting the value of the area ratio for each peak obtained in the same way from the measured pattern in this equation. In addition, a primary formula for determining the 1M illite content by the same method for two 1M unique peaks was also proposed [33]. This method was applied to the study of the determination of fault dating just after the study of van der Pluijm et al. (2001), applying IAA (Table 1 [3,5,6,7,21]). However, the quantitative values for each of the five peaks presented in this 2M1 polytype quantification method show significant differences. In particular, the hump appearing in the fine-size fraction with a high 1Md polytype content affects the setting of the intensity and width of other 2M1 and 1M peaks, which causes the error that the quantitative value is underestimated or overestimated.
The second method is a full-pattern-fitting method of simulated and measured patterns generated by WILDFIRE©. Ylagan et al. (2002) [34] developed a new code called PolyQuant, which is a quantification program automating the iterative matching process to find a ‘best fit’ between the mixed pattern of simulated 1Md and 2M1 patterns created in the forward modeling of WILDFIRE© and the measured pattern obtained from the size fractions. In particular, the optimal 1Md polytype simulated pattern selection process was automated by changing the crystallographic parameters. In this method, full-pattern-fitting was applied for the first time, and the difference was quantitatively presented by defining the objective function (J). In this respect, significant improvements have been made that are different from previous quantitative methods. Haines and van der Pluijm (2008) [8] proposed a least-squares lowest-variance approach based on WILDFIRE©, which is also essentially a full-pattern-fitting method, to find the best match between simulated and measured patterns (Table 1).
This WILDFIRE©-based polytype quantification method through full-pattern-fitting may seem to be theoretically the most proper quantification method that is likely to yield accurate results among the methods presented so far. Nevertheless, in the studies to which this method is applied, the error range of quantitative analysis values is significantly large (for example, [8,9,10,27,28,29]). The most fundamental cause for this is, above all, considered to be the limitation of the measured XRD pattern quality of the target sample obtained in the back-/side-packing state with conventional XRD equipment. In other words, it will be difficult to maintain consistency for each researcher in the randomness of the sample and the background correction in the XRD analysis procedure.
In order to overcome this problem, it is required to obtain (hkl) reflections by loading the sample into a capillary tube rather than back-/side-packing loading into an aluminum holder, and to perform background correction by measuring the empty tube. Minerals such as I-S (illite-smectite interstratified mineral) components may be contained in the sample, which may also be a hindrance to the process of finding the best-fit. The existence of I-S can be confirmed from the oriented XRD pattern in the low-angle 2θ range. Considering the fundamental particle concept, I-S can be viewed as 1 Md illite. Therefore, among the crystallographic parameters of WILDFIRE©, variables such as percentage of interlayered smectite and its hydration state, ordering of the illite/smectite (Reichweite), and K fraction in the structure can be used to reflect the effect of I-S in the simulated pattern [4].
Song et al. (2014) [14] obtained the measured pattern of the target sample obtained under optimized conditions such as using Micro-focused XRD equipment with a 2D-image plate attached and thin capillary tube (0.6 mm in diameter) for the first time. Reliable quantitative analysis results were obtained by iterative full-pattern-fitting this pattern with mixed patterns of 2M1 and 1Md illite based on WILDFIRE© made at various mixing ratios [14]. In addition, by presenting the R% value ((Σ|(simulated−measured)/simulated)|/n × 100) [36] similar to the objective function (J) suggested by Ylagan et al. (2002) [34], the degree of full-pattern-fitting was presented quantitatively. Since then [14,34], the R% value, or objective function (J) has been applied in a number of studies [13,14,15,16,19,20,22,23,26]. Figure 2 shows an example of polytype quantitative analysis of WILDFIRE©-based full-pattern-fitting.
In addition, methods using natural polytype end-member standards as synthetic mixtures without using WILDFIRE© [24,31] and methods such as BGMN®, Topas, Profex, AutoQuan, SIROQUANT, etc., based on Rietveld refinement [28] are also used. The Rietveld refinement method is basically a function of domain size, strain, and instrumental factors. Therefore, it does not consider the structural characteristics of clay crystallites, which is a factor of quantitative analysis error. Boles et al. (2018) [35] analyzed the results of evaluation experiments of WILDFIRE©-based full-pattern-fitting, End-member Standards Matching (STD) and Rietveld whole-pattern matching (BGMN®) quantitative analysis methods for synthetic samples. As a result, it was suggested that the WILDFIRE©-based full-pattern-fitting method is useful for samples with high 1Md illite content, in that it can utilize a data library of simulated patterns generated under various conditions of 1Md polytype [35].

6. Radiometric Dating Method

K-Ar and Ar-Ar methods have been applied as dating methods for size fractions separated from the fault gouge. Clauer et al. (2012) [37] discussed the limitations of the two methods through a comparative study. K-Ar dating requires a relatively large amount of sample (ca. 10–20 mg), although the amount required has reduced due to technological advancement, and since K content determination and Ar isotope analysis are separated in the analysis process, the analytical uncertainty is higher than that of Ar-Ar dating [37]. Therefore, duplicate or triplicate experiments are required in the K-content analysis process.
For fine size fraction, it is difficult to secure sufficient sample amount required for IAA application. Therefore, if dating is performed on a small sample, such as a fine fraction, Ar-Ar dating with low uncertainty of the result may be more appropriate. However, since the potential loss of Ar due to the recoil effect occurs in fine particles, there is a problem in that the encapsulation technique, which still requires technical verification, must be used to obtain reasonable results [3,5,6,8,9,10,11,12,17,18,21,25,31]. Therefore, for the fine size fraction, K-Ar dating may be more appropriate if the amount of size-separated sample is sufficient [37]. For this reason, in spite of recent developments in Ar-Ar dating, the conventional K-Ar method is still a valuable tool, because it is convenient and straightforward to use, with a high level of technical perfection and standardization. Therefore, the two dating methods can be used complementary to each other [37].
An additional point to be considered in dating is the presence of K-containing minerals, such as K-feldspar and biotite, that affect the dating results. In particular, the content of these minerals may be high at a relatively coarse particle size. In this case, it can be solved through Ar-release spectra analysis of Ar-Ar dating, but in most cases, it becomes an error factor in the dating result.

7. IAA (Illite-Age-Analysis) for Fault Dating

In the IAA (Illite-Age-Analysis) method, the first step is to graphically plot the dating data (y-axis) of three or more size fractions versus the relative content of 2M1 illite in each fraction (x-axis). From the simple linear extrapolation of the plots, the y-intercept value with a detrital 2M1 illite content of 0% is calculated. This y-intercept value is the generation age of 1M/1Md illite, that is, the fault activity age. Here, as the y-axis data, the value of exp(λt) − 1, which is a linear relationship with the radiogenic 40Ar/K ratio, rather than the age value, should be plotted against the relative content of 2M1 illite in each size fraction [1,2,3,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32].
The error of the fault dating result can be calculated from the value indicating the degree of fitting between the simulated pattern and the measured pattern in the polytype quantitative analysis process. The J value of Ylagan et al. (2002) [34] and the R% value of Song et al. (2014) [14] are values showing the degree of full-pattern-fitting. Song et al. (2014) [14] treated the R% value as the error range of the quantitative value determined for each fraction, and calculated the y-intercept value determined through its extrapolation as the error range of the 1Md illite generation age. In Figure 3, the IAA plot published in Song et al. (2014) [14] are presented as an example.
In addition, it is possible to confirm the reliability of the fault dating value by plotting the apparent K–Ar age value of each fraction against the illite crystallinity index (or Kübler index, defined as the half-height width (°∆2θ) of the illite (001) reflection of about 10 Å) [38], and by whether it is fitted with hyperbolic curves of negative correlations. In Figure 4, the K-Ar age value versus illite crystallinity index of each fraction published in Song et al. (2014) [14] are presented as examples.

8. Prerequisites and Procedures for Improvement of IAA

While reviewing the research results so far on fault dating by applying IAA, we introduced the details of the applied methods step by step, discussed the advantages and disadvantages, and suggested the prerequisites and process for obtaining reliable dating results. Table 2 summarizes the key prerequisites and suggestions mentioned above for each IAA step.
To summarize the important parts of the IAA, in step 1 of particle size separation, a pre-treatment process for exchangeable K+ removal and an appropriate particle size fraction were proposed. In the XRD analysis in step 2, the key is to obtain a measured XRD pattern including polytype (hkl) reflections by minimizing the preferred orientation for each fraction. For this purpose, the use of micro-focused XRD equipment with capillary tubes is strongly suggested. In the WILDFIRE©-based polytype quantification in step 3, full-pattern-fitting between an appropriate WILDFIRE©-based simulated and measured pattern is used as the most appropriate quantification method, but the quality of the measured pattern obtained in step 2 must be premised for reliability. High quantitative values can be obtained.
In the radiometric dating of step 4, both K-Ar and Ar-Ar methods are available, but it is a prerequisite to check whether K-containing minerals that can affect the age value are included. In the IAA of step 5, the value of exp(λt) − 1 should be plotted against the value of the relative content of 2M1 illite in each size fraction.

9. Concluding Remarks

The IAA method is a highly useful method to determine fault age, and there has been great progress in detailed methods for the last 20 years. In order to enhance the reliability of the IAA results, sub-processes should be performed in five steps.
Through a review of IAA applied studies, the prerequisites for each of the five steps (particle size separation process, XRD analysis process, polytype quantification, radiometric dating, IAA plot) and recommendations for improvement by detailed experimental content, etc., are summarized and presented. In particular, the application of micro-focused XRD with capillary tubes and improvement of the WILDFIRE©-based full-pattern-fitting method were strongly suggested as key to develop the reliability of IAA.
If the IAA method is improved as suggested in this study, it will be possible to determine the date of fault activity more reliably, and it will be an important tool for tectonic evolution research.

Author Contributions

Conceptualization, Y.S.; writing—original draft preparation, Y.S.; writing—review and editing, H.S. and Y.S.; supervision, Y.S.; project administration, Y.S.; funding acquisition, Y.S. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Korea Meteorological Administration Research and Development Program under Grant KMI 2018-02910.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pevear, D.R. Illite age analysis, a new tool for basin thermal history analysis. In Water-Rock Interaction; Kharaka, Y.K., Maest, A.S., Eds.; Balkema: Rotterdam, The Netherlands, 1992; pp. 1251–1254. [Google Scholar]
  2. Pevear, D.R. Illite and hydrocarbon exploration. Proc. Natl. Acad. Sci. USA 1999, 96, 3440–3446. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Van der Pluijm, B.A.; Hall, C.M.; Vrolijk, P.J.; Pevear, D.R.; Covey, M.C. The dating of shallow faults in the Earth’s crust. Nature 2001, 412, 172–175. [Google Scholar] [CrossRef] [Green Version]
  4. Reynolds, R.C., Jr. WILDFIRE: A Computer Program for the Calculation of Three-Dimensional X-ray Diffraction Patterns of Mica Polytypes and Their Disordered Variation; Wildfire©: Hanover, NH, USA, 1994. [Google Scholar]
  5. Solum, J.G.; van der Pluijm, B.A.; Peacor, D.R. Neocrystallization, fabrics and age of clay minerals from an exposure of the Moab Fault, Utah. J. Struct. Geol. 2005, 27, 1563–1576. [Google Scholar] [CrossRef]
  6. Van der Pluijm, B.A.; Vrolijk, P.J.; Pevear, D.R.; Hall, C.M.; Solum, J.G. Fault dating in the Canadian Rocky Mountains; Evidence for late Cretaceous and early Eocene orogenic pulse. Geology 2006, 34, 837–840. [Google Scholar] [CrossRef] [Green Version]
  7. Uysal, I.T.; Mutlu, H.; Altunel, E.; Karabacak, V.; Golding, S.D. Clay mineralogical and isotopic (K–Ar, δ18O, δD) constraints on the evolution of the North Anatolian Fault Zone, Turkey. Earth Planet. Sci. Lett. 2006, 243, 181–194. [Google Scholar] [CrossRef]
  8. Haines, S.H.; van der Pluijm, B.A. Clay quantification and Ar-Ar dating of synthetic and natural gouge: Application to the Miocene Sierra Mazatan detachment fault, Sonora, Mexico. J. Struct. Geol. 2008, 30, 525–538. [Google Scholar] [CrossRef]
  9. Haines, S.H.; van der Pluijm, B.A. Dating the detachment fault system of the Ruby Mountains, Nevada: Significance for the kinematics of low-angle normal faults. Tectonics 2010, 29, TC4028. [Google Scholar] [CrossRef]
  10. Schleicher, A.M.; van der Pluijm, B.A.; Warr, L.N. Nanocoatings of clay and creep of the San Andreas fault at Perkfield, California. Geology 2010, 38, 667–670. [Google Scholar] [CrossRef] [Green Version]
  11. Duvall, A.R.; Clark, M.K.; van der Pluijm, B.A.; Li, C. Direct dating of Eocene reverse faulting in northeastern Tibet using Ar-dating of fault clays and low-temperature thermochronometry. Earth Planet. Sci. Lett. 2011, 304, 520–526. [Google Scholar] [CrossRef]
  12. Rahl, J.M.; Haines, S.H.; van der Pluijm, B.A. Links between orogenic wedge deformation and erosional exhumation: Evidence from illite age analysis of fault rock and detrital thermochronology of syn-tectonic conglomerates in the Spanish Pyrenees. Earth Planet. Sci. Lett. 2011, 307, 180–190. [Google Scholar] [CrossRef]
  13. Chung, D.; Song, Y.; Park, C.-Y.; Kang, I.-M.; Choi, S.-J.; Khulganakhuu, C. Reactivated Timings of Some Major Faults in the Chugaryeong Fault Zone since the Cretaceous Period. Econ. Environ. Geol. 2014, 47, 29–38, (In Korean with English Abstract). [Google Scholar] [CrossRef] [Green Version]
  14. Song, Y.; Chung, D.; Choi, S.-J.; Kang, I.-M.; Park, C.; Itaya, T.; Yi, K. K-Ar illite dating to constrain multiple events in shallow crustal rocks: Implications for the Late Phanerozoic evolution of NE Asia. J. Asian Earth Sci. 2014, 95, 313–322. [Google Scholar] [CrossRef]
  15. Khulganakhuu, C.; Song, Y.; Chung, D.; Park, C.; Choi, S.-J.; Kang, I.-M.; Yi, K. Reactivated Timings of Inje Fault since the Mesozoic Era. Econ. Environ. Geol. 2015, 48, 41–49, (In Korean with English Abstract). [Google Scholar] [CrossRef] [Green Version]
  16. Bui, H.B.; Ngo, X.T.; Song, Y.; Itaya, T.; Yagi, K.; Khuong, T.H.; Nguyen, T.D. K-Ar Dating of Fault Gouges from the Red River Fault Zone of Vietnam. Acta Geol. Sin. (Engl. Ed.) 2016, 90, 1801–1811. [Google Scholar] [CrossRef]
  17. Fitz-Dıaz, E.; Hall, C.M.; van der Pluijm, B.A. XRD-based 40Ar/39Ar age correction for fine-grained illite, with application to folded carbonates in the Monterrey Salient (northern Mexico). Geochim. Cosmochim. Acta 2016, 181, 201–216. [Google Scholar] [CrossRef] [Green Version]
  18. Haines, S.; Lynch, E.; Mulch, A.; Valley, J.W.; van der Pluijm, B.A. Meteoric fluid infiltration in crustal-scale normal fault systems as indicated by d18O and d2H geochemistry and 40Ar/39Ar dating of neoformed clays in brittle fault rocks. Lithosphere 2016, 8, 587–600. [Google Scholar] [CrossRef]
  19. Song, Y.; Park, C.; Sim, H.; Choi, W.; Son, M.; Khulganakhuu, C. Reactivated Timings of Yangsan Fault in the Sangcheon-ri Area, Korea. Econ. Environ. Geol. 2016, 49, 97–104, (In Korean with English Abstract). [Google Scholar] [CrossRef]
  20. Bui, H.B.; Ngo, X.T.; Khuing, T.H.; Golonka, J.; Nguyen, T.D.; Song, Y.; Itaya, T.; Yagi, K. Episodes of brittle deformation within the Dien Bien Phu Fault zone, Vietnam: Evidence from K-Ar age dating of authigenic illite. Tectonophysics 2017, 695, 53–63. [Google Scholar] [CrossRef]
  21. Ring, U.; Uysal, I.T.; Glodny, J.; Cox, S.C.; Little, T.; Thomson, S.N.; Stübner, K.; Bozkaya, Ö. Fault-gouge dating in the Southern Alps, New Zealand. Tectonophysics 2017, 717, 321–338. [Google Scholar] [CrossRef]
  22. Sim, H.; Song, Y.; Son, M.; Park, C.; Choi, W.; Khulganakhuu, C. Reactivated Timings of Yangsan Fault in the Northern Pohang Area, Korea. Econ. Environ. Geol. 2017, 50, 97–104, (In Korean with English Abstract). [Google Scholar] [CrossRef]
  23. Jang, Y.; Kwon, S.; Song, Y.; Kim, S.W.; Kwon, Y.K.; Yi, K. Phanerozoic polyphase orogenies recorded in the northeastern Okcheon Belt, Korea from SHRIMP U-Pb detrital zircon and K-Ar illite geochronologies. J. Asian Earth Sci. 2018, 157, 198–217. [Google Scholar] [CrossRef]
  24. Kemp, S.J.; Gillespie, M.R.; Leslie, G.A.; Zwingmann, H.; Campbell, S.D.G. Clay mineral dating of displacement on the Sronlairig Fault: Implications for Mesozoic and Cenozoic tectonic evolution in northern Scotland. Clay Miner. 2019, 54, 181–196. [Google Scholar] [CrossRef]
  25. Lynch, E.A.; Mulch, A.; Yonkee, A.; van der Pluijm, B.A. Surface fluids in the evolving Sevier fold–thrust belt of ID–WY indicated by hydrogen isotopes in dated, authigenic clay minerals. Earth Planet. Sci. Lett. 2019, 513, 29–39. [Google Scholar] [CrossRef]
  26. Park, S.-I.; Noh, J.; Cheong, H.J.; Kwon, S.; Song, Y.; Kim, S.W.; Santosh, M. Inversion of two-phase extensional basin systems during subduction of the Paleo-Pacific Plate in the SW Korean Peninsula: Implication for the Mesozoic “Laramide-style” orogeny along East Asian continental margin. Geosci. Front. 2019, 10, 909–925. [Google Scholar] [CrossRef]
  27. Zhao, Q.; Yan, Y.; Tonai, S.; Tomioka, N.; Clift, P.D.; Hassan, M.H.A.; Aziz, J.H.B.A. A new K-Ar illite dating application to constrain the timing of subduction in West Sarawak, Borneo. GSA Bull. 2021. [Google Scholar] [CrossRef]
  28. Zwingmann, H.; Mancktelow, N.; Antognini, M.; Lucchini, R. Dating of shallow faults: New constraints from the AlpTransit tunnel site (Switzerland). Geology 2010, 38, 487–490. [Google Scholar] [CrossRef]
  29. Zwingmann, H.; Han, R.; Ree, J.-H. Cretaceous reactivation of the Deokpori Thrust, Taebaeksan Basin, South Korea, constrained by K-Ar dating of clayey fault gouge. Tectonics 2011, 30, TC5015. [Google Scholar] [CrossRef]
  30. Tonai, S.; Ito, S.; Hashimoto, Y.; Tamura, H.; Tomioka, N. Complete 40Ar resetting in an ultracataclasite by reactivation of a fossil seismogenic fault along the subducting plate interface in the Mugi Melange of the Shimanto accretionary complex, southwest Japan. J. Struct. Geol. 2016, 89, 19–29. [Google Scholar] [CrossRef]
  31. Aledga, L.; Viola, G.; Casas-Sainz, A.; Marcén, M.; Román-Berdiel, T.; van der Lelij, R. Unraveling multiple thermotectonic events accommodated by crustal-scale faults in northern Iberia, Spain: Insights from K-Ar dating of clay gouges. Tectonics 2019, 38, 3629–3651. [Google Scholar]
  32. Fisher, D.M.; Tonai, S.; Hashimoto, Y.; Tomioka, N.; Oakley, D. K-Ar dating of fossil seismogenic thrusts in the Shimanto Accretionary Complex, Southwest Japan. Tectonics 2019, 38, 3866–3880. [Google Scholar] [CrossRef]
  33. Grathoff, G.; Moore, D. Illite polytype quantification using WILDFIRE© calculated X-ray diffraction pat terns. Clays Clay Miner. 2016, 44, 835–842. [Google Scholar] [CrossRef]
  34. Ylagan, R.F.; Kim, C.S.; Pevear, D.R.; Vrolijk, P.J. Illite polytype quantification for accurate K-Ar age determination. Am. Mineral. 2002, 87, 1536–1545. [Google Scholar] [CrossRef]
  35. Boles, A.; Schleicher, A.M.; Solum, J.; van der Pluijm, B. Quantitative X-ray powder diffraction and the illite polytype analysis method for direct fault rock dating: A comparison of analytical techniques. Clays Clay Miner. 2018, 66, 220–232. [Google Scholar] [CrossRef]
  36. Chung, D.; Song, Y.; Kang, I.-M.; Park, C.-Y. Optimization of illite polytype quantification method. Econ. Environ. Geol. 2013, 46, 1–9, (In Korean with English Abstract). [Google Scholar] [CrossRef] [Green Version]
  37. Clauer, N.; Zwingmann, H.; Liewig, N.; Wendling, R. Comparative 40Ar–39Ar and K–Ar dating of illite-type clay minerals: A tentative explanation for age identities and differences. Earth-Sci. Rev. 2012, 115, 76–96. [Google Scholar] [CrossRef]
  38. Kübler, B. Les indicateurs des transformations physiques etchimiquesdans la diagènese, température et calorimétrie. In Thermométrie et Barométriegéologiques; Lagache, M., Ed.; French Society of Mineralogy and Crystallography: Paris, France, 1084; pp. 489–596. [Google Scholar]
Figure 1. Schematic diagram showing procedures for IAA with the conceptual IAA plot.
Figure 1. Schematic diagram showing procedures for IAA with the conceptual IAA plot.
Minerals 11 01162 g001
Figure 2. Example of polytype quantitative analysis of WILDFIRE©-based full-pattern-fitting. This figure was the same as Figure 14, published in Park et al. (2019) [26].
Figure 2. Example of polytype quantitative analysis of WILDFIRE©-based full-pattern-fitting. This figure was the same as Figure 14, published in Park et al. (2019) [26].
Minerals 11 01162 g002
Figure 3. Example of IAA plot for all size fractions of a fault clay samples. This IAA plot was the same as Figure 6, published in Song et al. (2014) [14].
Figure 3. Example of IAA plot for all size fractions of a fault clay samples. This IAA plot was the same as Figure 6, published in Song et al. (2014) [14].
Minerals 11 01162 g003
Figure 4. Example plots of the illite crystallinity index (or Kübler index) against the apparent K–Ar ages of all size fractions for a fault clay samples. Negative correlations were fitted by hyperbolic curves, converging to different ages. This plot was the same as Figure 9, published in Song et al. (2014) [14].
Figure 4. Example plots of the illite crystallinity index (or Kübler index) against the apparent K–Ar ages of all size fractions for a fault clay samples. Negative correlations were fitted by hyperbolic curves, converging to different ages. This plot was the same as Figure 9, published in Song et al. (2014) [14].
Minerals 11 01162 g004
Table 1. Summary of fault dating researches using IAA for last 20 years, in which fault names, selected size fractions, type of XRD equipment and holder, illite polytype quantification method, and raiometric dating method to each study result.
Table 1. Summary of fault dating researches using IAA for last 20 years, in which fault names, selected size fractions, type of XRD equipment and holder, illite polytype quantification method, and raiometric dating method to each study result.
No.Fault NameSize Fractions (µm)XRD Equipment with Sample HolderIllute Polytype QuantificationRadiometric DatingYearRef. No
1Lewis thrust<0.02, 0.02–0.2, 0.2–2ConventionalGrathoff and Moore (1996) method using WILDFIRE40Ar/39Ar20013
2Moab Fault, Utah<0.05, 0.05–0.5, 0.5–2ConventionalGrathoff and Moore (1996) method using WILDFIRE40Ar/39Ar20055
3Faults in Canadian Rocky Mountains<0.02, 0.02–0.2, 0.2–2ConventionalGrathoff and Moore (1996) method using WILDFIRE40Ar/39Ar20066
4Anatolian Fault<0.2, 0.2–0.5, 0.5–1, 1–2, >2ConventionalGrathoff and Moore (1996) method using WILDFIREK-Ar20067
5Sierra Mazatan detachment fault<0.05, 0.05–0.1, 0.1–0.5, 0.5–1, 1–2ConventionalLowest-variance approach using WILDFIRE40Ar/39Ar20088
6Fault of the Ruby Mountains<0.05, 0.05–0.4, 0.4–2ConventionalLowest-variance approach using WILDFIRE40Ar/39Ar20099
7San Andreas fault, Parkfield, Califonia<0.02, 0.02–0.2, 0.2–2ConventionalLowest-variance approach using WILDFIRE40Ar/39Ar201010
8Faults in AlpTransit deep tunnel site<0.1, 0.1–0.4, 0.4–2, 2–6, 6–10ConventionalSIROQUANT from Sietronics Pty Ltd.K-Ar201028
9West Qinling fault<0.05, 0.05–0.2, 0.2–2ConventionalLowest-variance approach using WILDFIRE40Ar/39Ar201111
10Pyrenean thrusts<0.05, 0.05–0.4, 0.4–2ConventionalLowest-variance approach using WILDFIRE40Ar/39Ar201112
11Deokpori Thrust<0.1, 0.1–0.4, 0.4–2, 2–6, 6–10Conventionalnot mentioned in detailK-Ar201129
12Chugaryeong fault zone, Korea<0.1, 0.1–0.4, 0.4–1, 1–2Micro-focused with capillary, 2D detectorIterative full-pattern-fitting with the WILDFIREK-Ar201413
13Daegwangri fault, Korea<0.1, 0.1–0.4, 0.4–1, 1–2Micro-focused with capillary, 2D detectorIterative full-pattern-fitting with the WILDFIREK-Ar201414
14Inje fault, Korea<0.1, 0.1–0.4, 0.4–1, 1–2Micro-focused with capillary, 2D detectorIterative full-pattern-fitting with the WILDFIREK-Ar201515
15Red River Fault, Vietnam<0.1, 0.1–0.4, 0.4–1, 1–2Micro-focused with capillary, 2D detectorIterative full-pattern-fitting with the WILDFIREK-Ar201616
16Mexican Fold-Thrust Belt<0.05, 0.05–0.2, 0.2–1, 1–2ConventionalLowest-variance approach using WILDFIRE40Ar/39Ar201617
17Faults in Death Valley and Panamint Valley<0.05, 0.05–0.2, 0.2–2ConventionalLowest-variance approach using WILDFIRE40Ar/39Ar201618
18Yangsan Fault in the Sangcheon-ri, Korea<0.1, 0.1–0.4, 0.4–1, 1–2Micro-focused with capillary, 2D detectorIterative full-pattern-fitting with the WILDFIREK-Ar201619
19Minami-Awa Fault<0.2, 0.2–0.5, 0.5–1, 1–2, 2–4X’Pert Pro Multi-purpose with capillaryIterative full-pattern-fitting with the WILDFIREK-Ar201630
20Dien Bien Phu Fault, Vietnam<0.1, 0.1–0.4, 0.4–1, 1–2Micro-focused with capillary, 2D detectorIterative full-pattern-fitting with the WILDFIREK-Ar201720
21Alpine Fault, New Zealand<0.1, 0.1–0.2, 0.2–0.5, 0.5–1ConventionalGrathoff and Moore (1996) method using WILDFIRE40Ar/39Ar201721
22Yangsan Fault in the Pohang Area, Korea<0.1, 0.1–0.4, 0.4–1, 1–2Micro-focused with capillary, 2D detectorIterative full-pattern-fitting with the WILDFIREK-Ar201722
23Faults in Yeongwol are, Korea<0.1, 0.1–0.4, 0.4–1, 1–2Micro-focused with capillary, 2D detectorIterative full-pattern-fitting with the WILDFIREK-Ar201823
24Río Grío, Vallès-Penedès Faults<0.1, 0.1–0.4, 0.4–2, 2–6, 6–10ConventionalIntegrated peak areas, using calibration constant for standard40Ar/39Ar201931
25Faults within Shimanto accretionary complex<0.2, 0.2–0.5, 0.5–1, 1–2, 2–4X’Pert Pro Multi-purpose with capillaryIterative full-pattern-fitting with the WILDFIREK-Ar201932
26Sronlairig Fault<0.05, 0.05–0.1, 0.1–0.2, 0.2–2X’Pert Pro Multi-purpose with capillarycorrected peak-area-measurement, Dalla Torre et al. (1994)K-Ar201924
27Sevier fold–thrust<0.05, 0.05–0.1, 0.1–0.5, 0.5–1, 1–2ConventionalLowest-variance approach using WILDFIRE40Ar/39Ar201925
28Faults in Chungnam Basin, Korea<0.1, 0.1–0.4, 0.4–1, 1–2Micro-focused with capillary, 2D detectorIterative full-pattern-fitting with the WILDFIREK-Ar201926
29Faults in West Sarawak, Borneo<0.2–0.5, 0.5–1, 1–2ConventionalIterative full-pattern-fitting with the WILDFIREK-Ar202127
Table 2. Summary of key prerequisites and suggestions for IAA method.
Table 2. Summary of key prerequisites and suggestions for IAA method.
StepDescriptions of ProceduresPrerequisites and Recommendations
1Particle size separation using high-speed centrifuge
-
Removal of exchangeable K+ ion by using dilute NaHCO3 for dispersion should be needed.
-
Flocculation by adding NaCl for extremely fine fraction should be required.
-
Removal of excess Na+ ions by using dialysis and freeze-drying should be required.
-
For particle sizes < 2 µm, 3 or more size fractions are required.
-
As the smallest size fraction, <0.1 µm size is recommended (ex. <0.02, <0.05, <0.1 µm).
2X-ray diffraction analysis for size fractions
-
Randomness of size fractions should be required for minimizing preferred orientation.
-
Using Micro-focused XRD with capillary tube is strongly recommended.
-
XRD pattern of empty capillary tube is required as background for capillary tube method
3WILDFIRE©-based polytype quantification
-
For conventional XRD and back/side-packing method, Grathoff and Moore (1996) [33] and internal standard method [35] seem to be appropriate.
-
For Micro-focused XRD with capillary tube, full-pattern-fitting method should be proper.
-
For quantification using full-pattern-fitting method, R% value can be determined as accuracy.
4Radiometric dating for size fractions
-
Both K-Ar and Ar-Ar methods are available.
-
For the Ar-Ar method, the vacuum-encapsulation method should be applied to prevent recoiling of 39Ar in the fine-size fraction.
-
Presence of K-containing minerals, such as K-feldspar, biotite, etc. should be checked.
5Illite-Age-Analysis (IAA) for dating fault age
-
Data should be plot exp(λt) − 1 vs. relative 2M1 content.
-
Plotting the age vs. illite crystallinity index (Kübler index) can be suggested for verification.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Song, Y.; Sim, H. Illite-Age-Analysis (IAA) for the Dating of Shallow Faults: Prerequisites and Procedures for Improvement. Minerals 2021, 11, 1162. https://doi.org/10.3390/min11111162

AMA Style

Song Y, Sim H. Illite-Age-Analysis (IAA) for the Dating of Shallow Faults: Prerequisites and Procedures for Improvement. Minerals. 2021; 11(11):1162. https://doi.org/10.3390/min11111162

Chicago/Turabian Style

Song, Yungoo, and Ho Sim. 2021. "Illite-Age-Analysis (IAA) for the Dating of Shallow Faults: Prerequisites and Procedures for Improvement" Minerals 11, no. 11: 1162. https://doi.org/10.3390/min11111162

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop