The following article is Open access

Mercury's Lobate Scarps Reveal that Polygonal Impact Craters Form on Contractional Structures

, , and

Published 2024 February 27 © 2024. The Author(s). Published by the American Astronomical Society.
, , Citation Chloe B. Beddingfield et al 2024 Planet. Sci. J. 5 52 DOI 10.3847/PSJ/ad1fff

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

2632-3338/5/2/52

Abstract

Analysis of polygonal impact craters (PICs) can be used to investigate the presence and orientations of subtle and/or buried faults and fractures across the solar system that may otherwise be unobservable in spacecraft images. Although this technique has been vetted for the analysis of extensional structures, no previous work has investigated if PICs also form on contractional thrust faults. This determination, which we investigated in this work, is critical for accurate tectonic setting interpretations from PICs. Mercury shows an abundance of thrust-fault-related landforms, making it an ideal laboratory to perform this investigation. In this work, we found that Mercury's thrust faults, and their overlying folds and fractures, cause some complex craters ∼20 km or larger to form PICs. However, in most cases, craters form as circular impact craters on these structures. When PIC straight rim segments do form, they parallel the lobate scarp thrust faults and fold hinges. Some PICs likely formed as a result of an impact's interaction with the thrust fault itself, while others may have interacted with fold hinge joints. The parallel relationship between PICs and shortening structures is consistent with the well-established relationship between PICs and extensional structures. Therefore, in addition to extensional fractures, contractional features should also be taken into consideration when utilizing PICs to interpret tectonic settings on bodies across the solar system.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

Polygonal impact craters (PICs), craters with one or more straight edges along their rims, have been identified on planetary bodies across the solar system (Figure 1) and form when an impact event occurs in a target with preexisting faults or fractures (Fielder 1961; Shoemaker 1962, 1963; Roddy 1978; Öhman et al. 2005; Beddingfield et al. 2016; Beddingfield & Cartwright 2020). Because PIC straight rim segments in many cases align with subimage resolution and/or subregolith fracture systems, they can provide crucial information for deciphering the tectonic histories of planetary bodies. For example, PICs have been used to more completely infer global-scale deformation patterns on the Saturnian moons Iapetus (Singer & McKinnon 2011) and Dione (e.g., Beddingfield et al. 2016) and the Uranian moon Miranda (Beddingfield & Cartwright 2020). However, only the association and relationship between PICs and extensional tectonic structures have been previously established. The association of PICs with contractional tectonic landforms has not been investigated on any planetary body or by utilizing physical models. Therefore, the interpretation of global-scale deformation patterns inferred from PICs is limited by the lack of knowledge regarding the possible association and orientation relationship between PICs and faults that accommodates lithospheric shortening.

Figure 1.

Figure 1. Examples of PICs on rocky and icy bodies across the solar system. (a) Mercury: Cunningham Crater (D = 37 km; coordinates: 30fdg385, 157fdg134), a PIC identified in this study within Caloris Basin. This figure shows a portion of MDIS image EW0220375119G. (b) Earth: Barringer (Meteor) Crater (D = 1.2 km; coordinates: 35fdg03, 111fdg02) in Arizona. This figure panel is a modified version of Figure 1(b) from Kumar & Kring (2008). (c) Mars: Martynov Crater, a PIC (D = 61 km; coordinates: 323fdg6, −30fdg4) identified by Öhman et al. (2006), located north of the Argyre basin. (d) Dione: a PIC (D = 27 km; coordinates: 20fdg1, 213fdg4) identified by Beddingfield et al. (2016) within the Wispy Terrain. (e) Iapetus: a PIC (D = 21 km; coordinates: −7°, 114°) identified by Singer & McKinnon (2011). (f) Miranda: a PIC (D = 11 km; coordinates: −36fdg4, 225fdg2) identified by Beddingfield & Cartwright (2020), located in the cratered terrain near Elsinore Corona.

Standard image High-resolution image

Mercury is known to exhibit PICs (e.g., Weihs et al. 2015), as well as craters superposed on shortening landforms, so-called lobate scarps and wrinkle ridges (e.g., Banks et al. 2015; Crane & Klimczak 2017). Both of these landform types are widely accepted to be the surface expressions of combinations of thrust faults and folds (Strom et al. 1975; Cordell & Strom 1977; Melosh & McKinnon 1988; Watters et al. 1998, 2001, 2004). Their formation is widely attributed to global contraction, the process causing Mercury to shrink from the long, sustained cooling of the planet (e.g., McKinnon 2014; Byrne 2018). Extensional landforms such as normal faults and graben are also curiously absent on Mercury's surface aside from those observed within plains units inside basins (Watters et al. 2009a, 2009b; Blair et al. 2013). Therefore, Mercury's surface represents an ideal test bed to investigate how impact cratering processes interact with preexisting contractional structures. In this work, we investigated if there is a relationship between PICs and lobate scarps to gain invaluable insight into how contractional structures affect crater morphologies.

Furthermore, the origin of PICs across the surface of Mercury remains uncertain because a general relationship between thrust faults and PICs has not been determined. Quantifying this relationship is therefore needed to use PICs as a tool to investigate Mercury's tectonic history. Detection and characterization of the orientations of subtle fracture systems would help discriminate between different tectonic processes on the innermost planet. For example, tidal despinning—the slowing of rotation to lock Mercury in its current 3:2 spin–orbit resonance with the Sun—is proposed to have formed a global fracture pattern (Klimczak et al. 2015); however, structural artifacts from this event and other early processes were likely overprinted by shortening landforms driven by global contraction, the volumetric reduction of Mercury due to long, sustained planetary cooling (Solomon 1977). Overprinting would have included the reactivation of favorably oriented fracture sets (Klimczak et al. 2015) but not the destruction of those unfavorably oriented. Those sets remain hidden but could be expressed during PIC formation. Therefore, characterization of PICs provides insight into how other tectonic processes, along with global contraction, played a role in shaping the surface of Mercury.

2. Background

2.1. Controls on Impact Crater Morphologies

Differences in impactor and target material properties influence the resulting impact crater morphologies. For a given impactor and impact velocity, the diameter of an impact crater will be larger on planets and satellites with lower gravity and lower target material density (e.g., De Pater & Lissauer 2010). Higher-velocity impacts will form craters with larger diameters, as will an increase in the density or size of the impactor. Impact crater geometries depend on other factors, including the angle of impact (e.g., Herrick & Forsberg-Taylor 2003), whether or not impacts are clustered (e.g., O'Keefe & Ahrens 1982; Schultz & Gault 1985b; Cochrane & Ghail 2006), the topography of the target area (e.g., Gifford & Maxwell 1979), the layering of the target material (Quaide & Oberbeck 1968), erosion (Ronca & Salisbury 1966), postimpact tectonic modification (e.g., Pappalardo & Collins 2005; Watters & Johnson 2010), and the presence of preexisting subvertical structures within the target material (e.g., Eppler et al. 1983; Kumar & Kring 2008). In addition to target material properties, crater geometries may also be affected by properties of the projectile including the porosity, composition, and shape (e.g., Schultz & Gault 1985a; Melosh 1989; Osinski & Pierazzo 2013).

Circular impact craters (CICs) are inferred to result from impact events in target material with uniform, multidirectional strength properties (e.g., Melosh 1989). This material could be nontectonized and so uniformly strong, or prefractured, if the fractures are widely or closely spaced relative to the size of the resulting crater (e.g., Fulmer & Roberts 1963). Additionally, CICs can form in a fractured target if the the fracture system is highly complex or covered by a thick layer of noncohesive sediment that limits interactions between the impactor and the underlying bedrock (e.g., Fulmer & Roberts 1963). In contrast, the only known cause for the formation of PICs is the presence of preexisting subvertical structures within the target material (e.g., Fielder 1965; Eppler et al. 1983; Öhman et al. 2005, 2008; Öhman 2009; Aittola et al. 2010). PICs exhibit straight rim segments, which reflect the orientations of preexisting fractures in the target material (e.g., Fielder 1961; Shoemaker 1962, 1963; Roddy 1978; Öhman et al. 2005). Consequently, CICs and PICs are excellent tools to distinguish between nontectonized and tectonized terrains on the surfaces of both silicate-rich and ice-rich surfaces across the solar system (e.g., Öhman et al. 2006; Beddingfield et al. 2016; Beddingfield & Cartwright 2020; Robbins & Riggs 2023).

2.2. Models of PIC Formation in Extensional Settings

Four PIC formation models have been investigated previously, each requiring the target material to contain preexisting fractures, which we refer to as models A, B, C, and D. Descriptions of these models are provided in detail by Öhman (2009), so we only summarize them here. According to model A, simple PICs are structurally controlled during the excavation stage of the transient crater (e.g., Schultz 1976; Eppler et al. 1983). The cavity expands in a direction oriented 45° to the surrounding fracture azimuths, forming PICs with azimuths that are offset by 45° to azimuths of the controlling fractures. This model is based on observations of two orthogonal fracture sets trending 45° to the straight crater rim segments of Barringer (Meteor) Crater located near Flagstaff, Arizona (Shoemaker 1963, 1977; Gault et al. 1974; Schultz 1976; Roddy 1978; Poelchau et al. 2008, 2009).

In model B, simple PIC shapes form as excavation flow preferentially overturns material along preexisting fractures, causing the crater to preferentially expand perpendicular to the fracture azimuths. In this model, the final PIC azimuths parallel surrounding fracture azimuths (Kumar & Kring 2008).

In model C, complex PIC geometries are determined during the modification stage (e.g., Schultz 1976; Eppler et al. 1983). In this model, the crater's straight rim segments are a result of the transient crater walls slumping via modification-related normal faulting along preexisting target structures along the crater wall. Consequently, the crater expands in a direction parallel to surrounding fracture azimuths. Like model B for simple PICs, this activity results in a parallel PIC–fracture relationship for complex craters.

In model D, applicable to both simple and complex craters, PICs inherit their geometries from movement of material along preexisting structures during the excavation stage (Öhman 2009). Like models B and C, model D predicts that the final PIC azimuths parallel surrounding fracture azimuths. Model D is supported by observational evidence of an association between faults and PIC crater rims on planetary surfaces (e.g., Gault et al. 1974; Reimold et al. 1998).

2.3. PICs throughout the Solar System

PICs have been identified throughout the solar system, and the relationships between PIC straight rim segment azimuths and controlling fracture azimuths have been investigated on many planetary bodies (e.g., Öhman 2009; Öhman et al. 2010). PICs are abundant on solid surfaces, and methods of automated detection of crater geometries are under investigation (Robbins & Riggs 2023). PIC azimuths on Mercury (Melosh & Dzurisin 1978; Strom & Sprague 2003) and Venus (Aittola et al. 2007, 2008, 2010; Öhman 2009) have been found to parallel azimuths of surrounding linear structures. On Earth, many PICs have also been identified, and their orientations have been compared to those of surrounding structures for both simple craters (e.g., Öhman 2009) and complex craters (e.g., Morrison 1984).

PICs have been used to infer global-scale deformation patterns on the Saturnian moons Iapetus (Singer & McKinnon 2011) and Dione (Beddingfield et al. 2016) and on the Uranian moon Miranda (Beddingfield & Cartwright 2020). Additionally, PICs have been investigated on Ceres (Zeilnhofer & Barlow 2021). Some PICs on Miranda overprint a terrain that has been interpreted to be made up of either contractional tectonic structures or cryovolcanically formed ridges (Schenk 1991; Beddingfield & Cartwright 2020). Earth's Moon also exhibits PICs (e.g., Fulmer & Roberts 1963; Melosh 1976; Schultz 1976; Eppler et al. 1983), and their azimuths parallel those of surrounding fractures (e.g., Melosh 1976; Schultz 1976; Eppler et al. 1983). Similarly, Martian PICs have been associated with the presence of preexisting target structures (Thomas & Allemand 1993; Watters 2006, 2009). PICs are also present on the surfaces of asteroids (Belton et al. 1994; Veverka et al. 1997; Thomas et al. 1999; Zuber et al. 2000; Prockter et al. 2002), the nucleus of a comet (Basilevsky & Keller 2006), and several other icy satellites (Smith et al. 1981; Plescia 1983; Porco et al. 2005; Helfenstein et al. 2005; Beddingfield et al. 2016, 2020). See Robbins & Riggs (2023) for a recent detailed summary of PICs identified throughout the solar system.

2.4. PICs on Mercury in a Contractional Tectonic Setting

Mercury exhibits PICs on its surface (Figure 1(a)) that were first systematically documented by Herrick et al. (2011) in the production of a global catalog of Mercurian craters. Weihs et al. (2015) visually inspected craters in this catalog and marked those with at least two straight rim segments as being polygonal, with 33 of the 291 assessed craters meeting this criterion. A recent study of PICs on Mercury systematically mapped and analyzed over 7000 impact craters, finding nearly 29,000 straight rim segments longer than 10 km (Yazici & Klimczak 2021; Yazici et al. 2024). These authors found that, in contrast to previous work that estimated PICs to represent ∼11% of craters on Mercury (Weihs et al. 2015), 83% of craters in their study contained straight rim segments. Both studies noticed roughly east–west orientations in straight rim segments near the poles; however, Yazici et al. (2024) also noted a weak preference for north–south orientations closer to the equator. Both studies located a range of sizes for PICs, 20–400 km (Yazici et al. 2024) and 65–240 km (Weihs et al. 2015).

Because PICs utilize preexisting fractures during their formation, the orientations of PIC walls in the spatial context of lobate scarps would inform our understanding of how folding progresses on Mercury. PIC straight rim segments aligned with the orientation of shortening landforms could indicate that pure shear accommodates folding and that folding and fracturing are coincident. Alternatively, PIC walls oriented obliquely to the fold hinge could be used to infer prefolding stress directions that may have generated tectonic fabrics prior to folding. It is also possible that both scenarios exist, and thus fracturing both before and during folding could be evaluated. The spatial density of PICs with along-trend versus oblique-to-trend walls could be evaluated to determine the spatial variability in tectonic fabric and shear strain. It is widely anticipated that many fracturing processes, unrelated to lobate scarp formation, occurred globally and regionally across Mercury's surface. Stresses resulting in these fractures may have been derived from tidal despinning (Klimczak et al. 2015), polar wander due to the Caloris Basin mascon (Matsuyama & Nimmo 2009), mantle upwelling or downwelling (King 2008; Tosi et al. 2013), or any combination of these processes over time. If sufficient in magnitude, stresses from these processes also would have produced fractures that resulted in PIC formation.

All models of PIC formation emphasize the necessity of subvertical fractures, regardless of their orientations. For lobate scarps, underlying major fault structures are predicted to dip between 5° (Galluzzi et al. 2019; Crane 2020) and 45° (Galluzzi et al. 2015; Watters 2021) on Mercury, with the theory of faulting predicting thrust faults to have optimal dips of 30° (Anderson 1951; Jaeger et al. 2009). The often rounded, asymmetric surface expressions of lobate scarps imply anticlinal folding above the underlying thrust faults. These folds on Earth are observed to contain fracture sets in various orientations relative to their limbs and hinges—not all of which are shallowly dipping (Figure 2; also see Figure 11 in Klimczak et al. 2019). Extension in the outer hinges causes near vertical shear fractures and jointing in these outer layers and shallowly dipping conjugate fracture sets in the inner hinges (Ramsay 1967; Cosgrove 2015). These joints may be accompanied by additional fractures caused by regional tectonic stresses, flexural slip, and intraunit stresses. Conjugate joint sets steeply dipping with orientations 60° from the hinge orientation may occur (Price & Cosgrove 1990) as well as nearly vertical joints perpendicular to the hinge (Reber et al. 2010), which may result from buckling (Eckert et al. 2014). Thus, horizontal compression and shortening, especially when in association with folding, do not limit the presence of steep fracture networks.

Figure 2.

Figure 2. Examples of anticline fractures that may result from folding above thrust faults on Mercury.

Standard image High-resolution image

Although Mercury's lobate scarps form from strong, horizontally compressive stresses, the crests of these landforms often display graben or opening mode fractures, signs of outer-arc extension, localized within the fold hinge (Byrne et al. 2018; Klimczak et al. 2019; Man et al. 2023). Where these fractures align with the overall trend of the landform, it may be interpreted that pure shear has dominated the strain history causing fracturing during folding (Ramsey 1968). However, fractures can also be observed crossing the hinge zone obliquely (Figure 2), implying that fractures accommodate regional stresses prior to folding (e.g., Ahmadhadi et al. 2008).

3. Data and Methods

We identified and investigated PICs on Mercury's lobate scarps by using a series of statistical tests. We utilized the R language to analyze rim azimuth distributions for each crater investigated. The steps discussed below are provided in additional detail in previous PIC studies (Beddingfield et al. 2016; Beddingfield & Cartwright 2020). For an overview of all steps outlined in the following subsections, see the flowchart in Figure 3. Due to the large number of craters analyzed in this study, the Appendix contains much of the information on crater and image data (Table A1), identified PICs (Figures A1A3), statistical test information (Table A2), collected azimuth data for PICs and lobate scarps (Figures A4A9), descriptions of lobate scarp proximities to craters (Table A3), and lobate scarp dip directions (Table A4).

Figure 3.

Figure 3. Flowchart illustrating our methodology.

Standard image High-resolution image

3.1. Crater Selection Criteria

We assessed the craters identified by Crane & Klimczak (2017) that are in overprinting stratigraphic relationships with lobate scarps. As summarized in Crane & Klimczak (2017), craters that are superposed by faults and those that overprint faults were distinguished. These identified craters are based on the thrust-fault-related data set provided by Byrne et al. (2014) and the Kinczyk et al. (2020) data set of craters with diameters ≥20 km. For each impact crater that matched our selection criteria, we determined the crater rim azimuth distribution.

For this study, our impact crater selection criteria included the following.

Criterion 1. The crater must overprint, and not be cut by, the lobate scarp. This stratigraphic relationship indicates that the lobate scarp existed before the impact event took place.

Criterion 2. The crater rims must not be cut or offset significantly by faults (unrelated to the lobate scarp). This criterion is based on the fact that postimpact-forming faults that cut across craters alter their plan-view morphologies (Galluzzi et al. 2015) and therefore may affect the results in later steps that involve accessing the rim azimuth distributions of these craters.

Criterion 3. For a similar reason to criterion 2, the crater rims must not be notably cut by overprinting large craters. The presence of these overprinting craters would be associated with large sections of missing rims of the crater in question, making analyses of rim azimuths challenging in later steps.

Criterion 4. The crater must be ∼20 km or larger in diameter. Because such a large number of craters are present on Mercury's lobate scarps, we are investigating craters of this size, which provide us with a large data set to analyze for purposes of this work. Analyses of smaller craters are beyond the scope of this project and will be assessed in future work.

3.2. Image Selection and Processing

All impact craters that meet our selection criteria were identified by utilizing the MErcury Surface, Space ENvironment, GEochemistry and Ranging (MESSENGER) Mercury Dual Imaging System (MDIS) Global Basemap (∼116 m pixel−1), also called the Morphology Mosaic, published in 2016 and available through the United States Geological Survey Astrogeology branch (Murchie et al. 2016; Hawkins et al. 2007; Denevi et al. 2016) and the Mercury Application of JMARS (Christensen et al. 2009). We then analyzed all individual MDIS images that cover each crater of interest to further investigate the craters that meet the selection criteria. Each crater used in this study was subsequently assigned a unique label for organizational purposes in this study (Appendix Table A1).

We utilized the highest-resolution MDIS images available that cover each selected impact crater overprinting a lobate scarp (Appendix Table A1). In some cases, multiple MDIS images were acquired for analyzing individual crater and lobate scarp sets. For example, multiple images were needed if a single image did not cover the entire crater, or if multiple images were needed to investigate the crater and a sufficient length of the adjacent lobate scarp. Additionally, in some cases multiple images were needed if different lighting conditions were more favorable for analyzing the lobate scarp than for the crater. All images used are detailed in the Appendix (Table A1). These MDIS images were downloaded from the Planetary Data System website. 4 We utilized these images for measurements using the Quantum Geographic Information System software (QGIS Development Team 2021).

Processing and projection of MDIS images were conducted using the Integrated Software for Imagers and Spectrometers 3, version 4.2.0 (Anderson et al. 2004). The images were associated with a camera model for MDIS and augmented with spatial information (geometries of the spacecraft, Sun angle geometries, ground positions, etc.). Because the geometries of surface features are most accurate, with negligible distortion, at the projection center, we projected each overlying MDIS image to the coordinates at the center of each impact crater using a sinusoidal projection. This technique allowed for high-accuracy azimuth measurements to be taken along the crater rims, which was critical for this analysis.

3.3. Crater Rim Tracing

Each impact crater rim was manually traced, as illustrated in Figure 4. We utilized shadowing and lighting variations as high-elevation indicators associated with the crater rims. The resultant crater polygons from this tracing step were then converted to sets of multilines by splitting the continuous polygon boundaries at their vertices. The numbers of vertices were determined based on the image resolutions (Table A1), where the distance between vertices is equivalent to the distance across a pixel. For each set of multilines, the azimuths of each individual segment were calculated. Each set of multiline segment azimuths and their associated lengths were exported to R for statistical analysis.

Figure 4.

Figure 4. Illustrations of the methodology for analyzing crater rim and nearby lobate scarp azimuths. Only the closest visible sections of each lobate scarp analyzed were investigated to better compare with PIC straight rim segments, as described in Section 3.4. (a) An example of the most common scenario in this study, where an impact crater (crater 3), uncut by extensional faults, other lobate scarps, or craters, overprints a single lobate scarp. (b) The traced crater (orange) and the section of the lobate scarp immediately adjacent to the crater (green) shown in panel (a). (c) Rose diagrams illustrating the azimuth distributions of the crater and lobate scarp traced in panel (b). (d) An uncommon example in this study, showing a crater (crater 1) that has visibly offset rim segments and is overprinting a lobate scarp. (e) The traced crater (orange) and the section of the lobate scarp immediately adjacent to the crater (green) from panel (d). (f) Rose diagrams illustrating the azimuth distributions of the crater and lobate scarp traced in panel (e). (g) An uncommon example in this study, showing a large crater (crater 2) with rim segments that are overprinted by smaller craters. Two nearby lobate scarps are equidistant to the large crater. (h) The traced crater (orange) and the sections of the two lobate scarps that are equidistant to the crater (green and cyan) shown in panel (g). (i) Rose diagrams illustrating the azimuth distributions of the crater and lobate scarps traced in panel (h).

Standard image High-resolution image

The crater rim tracing procedure was not affected by variations in image illumination angles. PICs are easier to recognize by eye in images with low illumination angles where prominent shadows are present along crater rims. However, studies have shown that, when rim azimuths are measured quantitatively, neither image resolution nor solar illumination effects due to lighting geometry have a strong effect on whether or not a crater is identified as a PIC (Binder & McCarthy 1972; Öhman et al. 2006). Measured rim azimuth distributions of impact craters taken on images with low illumination angles have been shown to be statistically similar to those taken on images with high illumination angles (Öhman et al. 2006).

None of the impact craters that we analyzed were notably cut by large extensional faults, lobate scarps, or large craters (see Section 3.1 above). In the majority of cases, the analyzed crater rims also were not cut by smaller features (Figure 4(a)). However, some craters that we analyzed were cut by small features, including faults or mass wasting features (Figure 4(b)) or small craters (Figure 4(c)). In these locations, the small sections of the crater rims erased by these features were excluded from our azimuth distributions (Figures 4(e) and (h)).

3.4. Lobate Scarp Tracing

Similar to our methodology for tracing crater rims, we also traced the structures associated with lobate scarps. We traced the thrust fault surface breaks exposed at the bases of the forelimbs, which are the steeper faces of the lobate scarps. We used shadowing and/or lighting of the surface topography as a guide. However, in some cases, identifying lobate scarp traces was more difficult than tracing impact crater rims due to their more shallow and subdued topography. In some images, shadows and variations in apparent surface brightness, caused by this more subdued topography, were less prominent and therefore more difficult to assess. Because the topography of lobate scarps includes a broad, gently dipping back limb and a steep forelimb, it was critical for us to utilize images with the appropriate lighting geometry.

Lobate scarps are mostly easily recognizable and traceable in images with lighting geometries that allow shadows to be cast along their steep sides, where the fault breaks the surface. Therefore, the direction of sunlight in an image must be from the direction of the shallow back limb in order for us to make use of this shadowing for purposes of tracing lobate scarps. To mitigate this issue, we investigated multiple MDIS images with different lighting geometries covering each lobate scarp (see Section 3.2 above). As mentioned in Section 3.2, multiple MDIS images were analyzed for the area covering each lobate scarp to obtain the images with the most ideal lighting geometry for each feature. In some cases, different images were used to trace the lobate scarp fault than what were used to trace the crater rim. The images determined to be the best for these measurements, and therefore utilized in this study, are listed in the Appendix (Table A1).

To compare lobate scarp azimuths to PIC azimuths in later steps, we investigated the segment of each lobate scarp immediately adjacent to the overprinting crater (Figure 4). Many lobate scarp segments that are immediately adjacent to the overprinting crater are covered by that crater's ejecta blanket (for example, the area south of the crater in Figure 4(a)). As a result, lobate scarp tracing was done along the segments that were not masked by ejecta and as close as possible to the crater (for example, the area north of the crater in Figure 4(a)). In the majority of cases, a single lobate scarp is associated with an analyzed crater (Figures 4(b), (c), (e), and (f)). However, in some cases, the measurable sections of two lobate scarps are equidistant from the crater (Figure 4(h)). In those cases, we incorporated both lobate scarps into our study, which we termed scarps A and B for each relevant study location (Figure 4(i)).

Like our crater rim tracing methodology, the resultant traced lobate scarp lines were converted to multiline sets, and the azimuths of each segment were calculated. For each lobate scarp, the set of azimuths and their associated lengths were exported to R for statistical analyses.

3.5. Chi-squared Tests

To identify PICs, we tested for uniform azimuth distributions for each analyzed impact crater by applying Pearson's chi-squared tests (e.g., Burt et al. 2009). For these tests, we selected alpha levels of 0.05. Therefore, if a resulting p-value of the Pearson's chi-squared test is less than the alpha level of 0.05, then there is 95% confidence that the data are not uniformly distributed. For each impact crater trace, the set of multiline segment azimuths and lengths were utilized to test for a uniform distribution of crater rim azimuths, normalized for the lengths of each measurement, using R's chisq.test function.

We binned each crater rim azimuth data into 8° and 16° bins. We then applied four chi-squared tests to each crater, shifting the data within the bins by 4° and 8°, respectively. This shifting method allowed us to detect PIC straight rim segments more accurately in the cases where the orientations of these segments fall close to a bin threshold and therefore may otherwise be split across two bins. This methodology also allowed us to account for straight rim segments of different lengths. If the result of a Pearson's chi-squared test supported the null hypothesis, the azimuth distribution of that particular crater was considered to be uniform. In these scenarios, the crater was identified as a CIC. Alternatively, if the test result was significant, then the azimuth distribution was not considered to be uniform, and the analyzed crater was identified as a PIC.

3.6. Comparing PICs with Lobate Scarps

The prominent PIC rim azimuths (also referred to as "straight rim segments") were determined for all identified PICs. The prominent rim azimuth(s) of each PIC is (are) reflected by the mode(s) of the azimuth distributions. The modes for each PIC were determined using R's dip.test function (Maechler & Ringach 2013), which applies the dip test of unimodality described by Hartigan et al. (1985). The modes of each lobate scarp azimuth distribution were also determined using this method.

We then compared the prominent PIC rim azimuth(s) with those of adjacent lobate scarps. We did not expect PIC straight rim segments to precisely reflect the orientations of the measured lobate scarp sections, even if the relationship between PICs and lobate scarps is truly parallel. We had this expectation because the lobate scarps were traced at some distance from the overprinting PICs (see Section 3.4), and the orientations of the lobate scarps can vary slightly across short distances (see the lobate scarps in Figures 4(a)–(h)). We accounted for this expected azimuth spatial variation by allowing for an acceptable range of differences in PIC and lobate scarp straight segment azimuths. As a threshold, we considered PIC straight rim segments to subparallel lobate scarps if the differences between their prominent azimuths fell into a category within five bins or less (out of 60 bins). Because each bin contains 6° of azimuths, this threshold is equivalent to <30°.

4. Results and Discussion

4.1. PICs on Lobate Scarps

As shown in Table 1, 29 PICs overprinting lobate scarps were identified out of the 163 craters analyzed in this study. Therefore, our results indicate that PICs can form on shortening landforms, in some cases. However, because <20% of the analyzed craters are PICs, we find that CIC formation is the most common outcome of impact events on shortening landforms. Of the identified PICs, 21 have one straight rim segment, while eight exhibit two straight segments.

Table 1. Results Showing the Modes of PIC and Lobate Scarp Azimuths

Crater IDPIC ModesLobate Scarp ModesParallel?Azimuth Difference
 1st2ndAB DegreesNo. of Bins
136°–42° 48°–54° Yes6°–12°2
278°–84°174°–180°24°–30°54°–60°Yes18°–24°4
336°–42° 144°–150° No  
40°–6° 12°–18°60°–66°Yes6°–12°2
11162°–168° 168°–174° Yes0°–6°1
130°–6°114°–120°174°–180° Yes0°–6°1
156°–12°84°–90°18°–24° Yes6°–12°1
16132°–138°42°–48°6°–12°162°–168°Yes24°–30°5
17126°–132°78°–84°66°–72° Yes6°–12°1
180°–6° 54°–60° No  
2166°–72° 0°–6°48°–54°Yes12°–18°3
2278°–84°138°–144°0°–6° No  
3366°–72° 0°–6° No  
3460°–66° 30°–36° Yes24°–26°5
386°–12° 120°–126° No  
49144°–150° 168°–174°114°–120°Yes18°–24°4
5130°–36°126°–132°144°–150° Yes12°–18°3
76168°–174° 18°–24° Yes24°–30°5
8454°–60°138°–144°174°–180°36°–42°Yes12°–24°3
102114°–120° 120°–126°42°–48°Yes0°–6°1
115126°–132° 24°–30° No  
116120°–126° 6°–12°54°–60°No  
12266°–72° 36°–42°156°–162°Yes24°–30°5
134144°–150° 138°–144° Yes0°–6°1
1400°–6° 138°–144°24°–30°Yes18°–24°4
145138°–144° 132°–138° Yes0°–6°1
148132°–138° 114°–120°72°–78°Yes12°–18°3
152120°–126° 12°–18° No  
15978°–84° 36°–42° No  

Note. All craters that have at least one straight rim segment are shown. Information on all 163 craters analyzed is provided in the Appendix.

Download table as:  ASCIITypeset image

We find that many PIC straight rim segments exhibit parallel relationships with adjacent lobate scarps. Of the 29 PICs identified in this study, 20 exhibit a parallel relationship with an adjacent lobate scarp. Four examples are shown in Figure 5. See the information provided in the Appendix for additional details on other PICs that parallel adjacent lobate scarps (Figures A4A9). As summarized in Table 1, five of the identified PICs exhibit azimuths between 0° and 6° of an adjacent lobate scarp. Four PICs exhibit azimuths between 6° and 12°, four between 12° and 18°, three between 18° and 24°, and four between 24° and 30°.

Figure 5.

Figure 5. Example illustrations of our results, showing that PICs with straight rim segments parallel adjacent lobate scarps. (a) Crater 13. (b) Crater 140. (c) Crater 134. (d) Crater 4. For a full illustration of our results for all PICs identified, see the annotated rose diagrams in Figures A4A9. For images of all PICs identified, see Figures A1A3. See Table A1 for coordinate and image information. See Table 1 for information on azimuth modes, indicating straight segments for all PICs and adjacent lobate scarps analyzed.

Standard image High-resolution image

4.2. CICs on Lobate Scarps

Most impact events that occur on lobate scarps do not form as PICs and instead were identified as CICs. There are multiple possible explanations for this result. Perhaps the craters that form as CICs did so because of one or more of the following reasons. (1) The impact event occurred too far away from the fault surface break and/or folding-associated fractures. In this case, these structures may not be close enough to influence the orientation of the resulting rim during crater formation. Or, few to no fracture sets are present in the anticline overlying the thrust fault. (2) The craters formed too directly on top of the thrust fault. In this case, the associated thrust and many of the associated steep fractures would be oriented perpendicularly to the crater rim, which is not optimal for influencing rim development. Additionally, large portions of the crater rims may have formed too far from the lobate scarp in this scenario and therefore were not affected by the underlying fault and fractures. (3) The craters are too large relative to the lobate scarp, such that they excavated into the footwall of the thrust. (4) The craters are too small relative to the lobate scarp and therefore did not interact with the underlying thrust fault or fractures. (5) Our methodology is conservative; therefore, some straight rim segments may not have been identified. For example, straight rim segments and PICs may not have been identified if the crater is highly degraded, the straight segment does not take up a large enough portion of the crater rim, or the MDIS image resolution is low relative to the crater size.

We investigated the above possible explanations for CICs on lobate scarps by comparing the study location characteristics of identified PICs with those of CICs (Table 2, Figure 7). We did not find a notable relationship between crater diameter and crater shape (Figure 6). Identified PICs range in diameter from ∼20 to 81 km, with the average being 37 km. In comparison, diameters of CICs range from ∼20 to 138 km with an average of 43 km.

Figure 6.

Figure 6. Box-and-whisker plots of PIC and CIC diameters. See Tables A1 and A2 for crater diameters and classifications, respectively. The crosses in the blue boxes represent the means, and the middle horizontal lines represent the modes. The top and bottom lines of the boxes represent the medians of the third and first quartiles, respectively. The ends of the vertical lines represent the maximum and minimum values.

Standard image High-resolution image

Table 2. Comparison Summary of PIC and CIC Characteristics

DescriptionPICsCICs
DiameterMinimum20 km18 km
 Average37 km43 km
 Maximum81 km138 km
 
Lobate scarp–crater intersectionYes, center41%62%
 Yes, edge41%10%
 None31%43%

Note. Values are provided that show the percentage of PICs associated with lobate scarps that run approximately through the crater center and crater edge and outside the crater edge (none). See Figure 7 for an illustration of these geometries. The percentages along each column, for PICs and CICs, add up to over 100% because some craters are associated with multiple lobate scarps. For additional information on these descriptions, see the caption of Figure 7. Also, see Appendix Table A3 for the data used to derive these results.

Download table as:  ASCIITypeset image

We also investigated more specific relationships between craters and adjacent lobate scarps (Figure 7; Tables A3 and A4). We categorized lobate scarps with a surface thrust fault that (1) intersects the crater through the crater center, (2) intersects the crater along the crater edge, and (3) does not intersect the crater. In the latter case, while the crater overprints the lobate scarp, the associated surface thrust fault does not directly underlie the crater. Therefore, the crater in question is instead overprinting a subsurface thrust fault in these locations.

Figure 7.

Figure 7. Illustration of crater and surface thrust fault configurations. Also see Table A3. The surface thrust faults associated with the lobate scarps are shown in gray. The values indicate the percentage of particular configurations between the craters and lobate scarp surface thrust faults (center, edge, or none) with the identified crater geometry (PIC vs. CIC). The percentages for PICs and CICs each total more than 100% because some craters were associated with multiple lobate scarps. Our results indicate that, when an impact event occurs on top of a surface thrust fault, the resulting crater will be more likely to form a PIC than a CIC if the center of the impact is offset from the surface thrust fault (see red crater). In this scenario, 41% of PICs identified exhibit this configuration with at least one lobate scarp surface thrust fault. In comparison, only 10% of CICs exhibit this configuration. In Figure 5, three of the four examples shown of PICs overprinting lobate scarps fall into this "edge" category (panels (a), (b), and (d)), while only one example falls into the "center" category (panel (c)).

Standard image High-resolution image

We found a small difference in percentage of PICs and CICs that are associated with crater center intersections with lobate scarp surface thrust faults. These cases are associated with 41% of identified PICs and 62% of identified CICs. We also found only a small difference in the percentage of PICs and CICs associated with nearby surface thrust faults but without direct intersections with these faults. These configurations are associated with 31% of PICs and 43% of CICs. We found a notable difference in crater geometries that have edge intersections with surface thrust faults. In these scenarios, 41% of identified PICs are associated with at least one lobate scarp surface thrust fault that skims the crater edge, while only 10% of CICs exhibit this configuration. Therefore, our results indicate that in the presence of shortening structures, PICs are most likely to form when the impact event occurs so that the resulting crater center is offset from the lobate scarp surface thrust fault.

However, we also find that PICs form in other configurations (center and none in Figure 7). Perhaps PICs on structures accommodating shortening form when an impact event causes the crater walls to interact with the shallow subsurface portion of thrust faults. Alternatively, perhaps these PICs form as a result of interactions with secondary structures instead of the main thrust fault. For example, perhaps the presence of extensional fractures, thought to form in response to lobate scarp formation (Figure 2; e.g., Engelder & Geiser 1980; Klimczak et al. 2019), resulted in some impact events occurring in these contractional terrains to form as PICs. In other words, on shortening landforms, PICs may be most likely to form when the crater edge overprints a surface thrust fault and/or overprints the fractures associated with the thrust's overlying fold.

4.3. Implications for PICs and Lobate Scarps

Our results indicate that impact events commonly produce PICs when the impact center location is slightly offset from the lobate scarp surface thrust fault (see red PIC illustration in Figure 7). This geometry allows the resulting crater rim to interact with the fault.

Some past studies conclude that complex PICs form from normal faulting during the modification stage of crater formation (model C in Section 2.2; e.g., Schultz 1976; Eppler et al. 1983), while other studies instead conclude that they form from thrust faulting during the excavation stage (model D in Section 2.2; Öhman 2009; see Section 2.2). Öhman (2009) makes the argument that the fact that the straight rim segments of PICs are topographic highs indicates that thrusting is necessary, thereby supporting model D.

However, in this work, we note an example of a PIC where crater wall slumping appears to be associated with the formation of an impact crater straight rim segment. In this example, one part of the crater rim appears to have partially formed a straight rim segment but exhibits "failed" slumping (Figure 8). Perhaps this crater represents an example of an intermediate step of PIC straight rim segment formation during the crater modification phase. This slumping may be the result of backsliding against the original thrust motion of the fault plane due to the back limb of the lobate scarp being unsupported after crater formation, similar to slip sheets and collapse folds observed on Earth (Perucca et al. 2016; Harrison & Falcon 1934, 1936). These observations support model C, that modification stage normal faulting forms complex PIC straight rim segments. Based on our observations coupled with the logic described by Öhman (2009), we make the interpretation that both the excavation (model D) and modification (model C) stages of complex PIC crater formation contribute to the formation of the straight rim segments in the presence of thrust faults.

Figure 8.

Figure 8. Example of a crater with evidence for wall slumping associated with a straight crater rim segment. Here we show crater 122 (D = 34 km) in different MDIS images with differing lighting geometries. Perhaps this crater represents an example of an intermediate step of PIC straight rim segment formation during the crater modification phase, which would support model C for PIC formation (e.g., Schultz 1976; Eppler et al. 1983; see Section 2.2) for PICs associated with thrust faults. MESSENGER MDIS images (a) EN0227933432M, (b) EN0245441590M, and (c) EN0219610891M.

Standard image High-resolution image

As described in Section 4.3, perhaps the presence of joints associated with folding during lobate scarp formation explains the formation of PICs in other configurations ("center" and "none" categories illustrated in Figure 7). These PICs, which commonly show parallel relationships with adjacent lobate scarps, form as a result of the existence of small joints that trend parallel to the underling thrust fault or fold hinge. Additionally, perhaps the parallel relationships in these scenarios between PICs and lobate scarps indicate that joints related to anticline formation more often form parallel than obliquely to the fold axes of Mercury's lobate scarps.

Folding in the brittle parts of the lithosphere is widely known to be accommodated by fractures (e.g., Engelder & Geiser 1980), and many fracture orientations within folds exist (e.g., Klimczak et al. 2019). Therefore, perhaps these particular PICs are highlighting the internal structural architecture of these thrust-fault-related landforms. Perhaps PIC straight rim segments that are aligned with lobate scarp orientations indicate that pure shear accommodates folding and that folding, fracturing, and faulting are coincident within lobate scarps on Mercury.

Perhaps PICs with straight rim segments that are oriented obliquely to adjacent lobate scarps are reflecting fractures unrelated to the lobate scarp. Perhaps they are instead reflecting prefolding and/or postfolding stress directions that generated tectonic fabrics prior to or following lobate scarp formation. Alternatively, these PICs may be reflecting joints associated with anticline formation but with orientations that formed obliquely to the underlying thrust fault. In future studies, the locations and orientations of PICs with straight rim segments that are oblique to the lobate scarp could be evaluated to investigate the orientations of additional stress directions in these locations.

4.4. Comparison with Extensional Settings

Similar to our results for contractional settings, both CICs and PICs form in targets that exhibit extensional fractures and faults. There are many known causes for the formation of CICs in extensional settings, which have been noted on many planetary bodies and in physical laboratory experiments. For example, within the pervasively fractured Wispy Terrain on Dione, 76% of the impact craters analyzed were classified as PICs, while 24% were classified as CICs. In Dione's more subtly fractured "Non-Wispy Terrain," percentages of PICs are as little as 20% in some locations, where fractures were inferred (Beddingfield et al. 2016). Within Miranda's cratered terrain, which exhibits a large number of fractures of various sizes, only 29% of the identified craters were identified as PICs (Beddingfield & Cartwright 2020). However, the locations of fractures in these terrains relative to each crater analyzed are not as well constrained as those for Dione's Wispy Terrain. Therefore, the true percentage of PICs overprinting extensional features may be higher.

Many characteristics of prefractured target material have been attributed to the formation of CICs. As summarized in Fulmer & Roberts (1963), CICs are shown to form in fractured material if the target consists of a complex set of closely spaced fractures, very widely spaced fractures, or unconsolidated material (Fulmer & Roberts 1963). For example, joints and normal faults both form in response to extensional stresses on planetary bodies. If joints or normal faults are covered by a thick layer of regolith, comparable to or greater than the depth of the impact crater, then we would expect an impact event to form a CIC instead of a PIC.

CICs may also be more likely to form if the impact event creates a crater that is too large or too small relative to the fracture sizes and/or spacing (see Öhman et al. 2005 for a full summary). As summarized in Öhman et al. (2005), in many extensional settings, PICs are somewhat constrained to specific crater diameter ranges. For example, Schultz (1976) concluded that lunar PICs more often exhibit diameters of >1 and <15 km. Similarly, in the Argyre region on Mars, the majority of identified PICs fall within the 10–35 km diameter size range (Öhman et al. 2006).

The explanations for the presence of overprinting CICs along with PICs in extensional target material may provide explanations for the abundant CICs that we identified in contractional settings this work. If similar relationships between CICs and tectonic structures hold true for thrust faulting, then our finding of a large number of CICs relative to PICs is not surprising. In the case of lobate scarps on Mercury, thrust faults are more widely spaced than many of the extensional terrains on planetary bodies noted above. Therefore, we would expect a lower ratio of PICs relative to CICs on lobate scarps relative to extensional terrains. However, additional studies that investigate the relationships between CICs and the characteristics of contractional terrains are needed.

4.5. Applications for Future PIC Interpretations

Our results show that, in addition to extensional structures, contractional structures should be considered as possible explanations for the presence of PICs on bodies across the solar system. Our new knowledge of the parallel relationship between PICs and contractional structures can be applied to better interpret tectonic settings and global stress mechanisms on rocky and icy bodies where PICs are present, as described in Section 4.4.

For example, future analyses of other PICs elsewhere on Mercury could be used to obtain a more complete understanding of the extent and orientations of tectonic features, including the possible existence of contractional structures, that are difficult to discern in available spacecraft images. Detection and characterization of the orientations of subtle tectonic systems could help further discriminate between different tectonic processes. For example, tidal despinning—the slowing of rotation to lock Mercury in its current 3:2 spin–orbit resonance with the Sun—is proposed to have formed a global fracture pattern (Klimczak et al. 2015). If tidal despinning was overprinted by global contraction—the volumetric reduction of Mercury due to long, sustained planetary cooling (e.g., Solomon 1977)—it would utilize and reactivate favorably oriented fracture sets (Klimczak et al. 2015). Therefore, characterization of additional PICs in future work would provide insight into how other tectonic processes, along with global contraction, played a role in shaping the surface of Mercury. Future analyses may also provide insight into the tectonic setting related to Caloris Basin.

Consideration of contractional features in future PIC work could lead to more refined interpretations of global stress mechanisms on many icy bodies where PICs have been identified (see Section 2.3). For example, global stress events such as orbital recession, despinning, volume contraction, nonsynchronous rotation, and true polar wander may create regions of compression in some locations and tension in others (e.g., Collins et al. 2009). However, only extensional tectonic structures associated with regions of tension have been considered when interpreting PICs. For example, previous analyses of PIC locations and orientations on Dione yielded an inferred fracture pattern across the surface that points to satellite despinning and volume expansion (Beddingfield et al. 2016). Using our results, new analyses of these craters could be used to better interpret PICs in regions that would be in compression during these events (see stress maps provided by Collins et al. 2009 and references within).

5. Conclusions

We conclude that the presence of thrust faults and their associated folds and fractures can result in the formation of PICs for complex craters ∼20 km in diameter or larger. Additionally, we conclude that PIC straight rim segment azimuths exhibit parallel relationships with the controlling thrust fault azimuths, which are also parallel to their overlying fold hinges. These relationships between craters and shortening structures are like the relationships between craters and normal faults and joints observed in laboratory work and on solid planetary surfaces of rocky and icy bodies.

While PICs do form on thrust faults, we find that in most cases CICs form on thrust faults. In extensional settings, CICs are also known to form when overprinting and/or adjacent to normal faults and joints. However, the specific occurrences of PICs relative to CICs in extensional settings is not well constrained; therefore, we are not able to determine if the ratios of PICs to CICs on thrust faults are like those formed by normal faults. Additionally, only some of the PIC straight rim segments identified in this work may have formed as a result of the presence of the lobate scarp thrust fault itself. In addition to the thrust fault, some PICs may have formed in response to interactions with joints related to the formation of the lobate scarp fold. Future work is needed to further investigate this possibility and therefore determine if some PICs overlying lobate scarps can be utilized to better constrain fractured rock masses associated with lobate scarps. Additional work is also needed to determine if the relationship between PICs and lobate scarps holds true for craters with diameters smaller than 20 km.

Acknowledgments

This work was funded by the NASA Discovery Data Analysis Program, grant No. 80NSSC21K1016.

Appendix

Here we provide information regarding the locations of all craters analyzed in this study and the MESSENGER MDIS images used to analyze each crater (Table A1). See Section 3.1 for information on and justification of our crater selection criteria. Here we also provide all data and results for this work. Results of chi-squared tests (p-values) are provided (Table A2). See Section 3.5 for the description of how these tests were used to identify PICs. Results describing the proximity of lobate scarps to craters are provided in Table A3, results describing lobate scarp dip directions are provided in Table A4.

Table A1. IDs, Coordinates, Diameters, and Details on MDIS Images Used for Each Impact Crater Analyzed in This Work

CraterLatitudeLongitudeDiameterMDISImage
ID(deg)(deg)(km)Image(s)Resolution(s)
     (m pixel−1)
159.154129.32921EW0215937434G193
254.562−2.03922EW1017441552G148
356.751−27.34322EW0228154808G122
458.88527.43023EW0252639343G154
566.00443.24041EW0219903748G143
660.15550.58626EW0219776192G167
767.14061.53525EW0219648898G135
852.57268.74927EW0219436182G202
955.19759.29931EW0219606184G187
1054.28151.85830EW0234791785G159
1164.98871.19840EW1028963152D169
1266.57572.45020EW0254914294G198
1350.73593.65132EW0261828115G188
1455.683123.62734EW0231135611G175
1562.359133.49120EW1000244400G201
1653.737157.10067EW0243997316G220
1758.807160.41067EW0218204261I170
1850.933177.26141EW0260704693G112
1946.62192.11932EN1046571115M32
    EW0234240064G186
2040.93083.94745EW0249585137G233
2148.01564.84547EW0234537072G189
2248.85245.47126EW0255202026I152
2339.28131.91080EW0235088520G227
2448.90713.12645EW0250852619G175
2541.70818.16240EW0252870070G223
2643.4139.47477EW0265773054G158
2745.44517.00345EW0250506911G158
2846.953−30.59941EW1005255605I142
    EW0213025742G140
2945.368−28.87727EW0212982293G143
3044.667−27.44048EW0225863816G189
    EW0213025827G154
3140.401−137.95635EW0247626443I182
    EW0242881468G197
3250.322−147.84524EW0260215030G111
    EW0257938432G299
    EW0257938432G299
3347.375−147.51424EW1021876602B76
    EW1021876602B161
3449.778−178.17874EW0250392934G197
3543.703−178.31943EW0228024056G204
3644.394−175.70326EW0228024056G127
    EW0228024056G209
3729.548−169.24327EW0248000595I262
    EN0212676209M111
    EN0212676120M118
3836.219−164.45338EW0247943200G211
3934.840−162.73127EW0247914368G214
    EN0212502490M102
4027.216−162.41524EW0247885442G243
    EN0212502235M122
4137.292−122.95128EW0257563816G181
4231.976−104.21725EN0211676639M113
4334.034−50.16926EW0213373594G178
4429.847−46.94328EW0213330191G197
4529.185−20.39245EW1005111302G193
    EW0228155201G219
4632.848−18.85039EW0212808634G192
4727.800−8.95254EW0212678355G233
4837.976−7.18635EW1002232278G159
4934.376−0.37435EN0220762110M105
    EW0242841168G257
5034.6617.30030EW0235555178G248
5131.762−14.67846EN0235850479M119
5232.83312.51230EW0235342993G225
5338.43378.78524EN0239119260M52
5433.247105.76324EW0231395066G119
5538.300103.70225EW0246562104G220
5634.883104.98476EW1046195252G232
5734.647132.72542EW0220936688G89
5830.027137.02359EW0248778334G256
    EW0220850326G95
    EW0220807126G96
5923.402145.05928EN1014993813M49
6025.18478.41878EW0254654238G252
    EN0219222948M120
6119.66681.18124EW1003613015G259
6218.93253.87226EN0219647602M135
6318.1469.14330EN0220502267M74
6424.470−44.36330EW0213330292G226
    EW0243710758G279
6525.023−93.61733EN1006291928M45
    EN0226710114M98
6625.944−91.85133EN0211415553M140
    EN0226710146M95
6718.421−165.70424EN0212501849M153
6812.420−176.02047EN0258224295M102
    EN0212849080M190
6912.080−173.78335EN0212849200M182
709.277−169.82045EN0260821552M153
717.491−125.62827EN0257562934M111
    EN0227048246M148
723.715−110.35734EN0226921108M155
737.857−78.11125EN0226455550M139
    EW0213852099G174
744.2835.28535EW1045210266H164
7510.191−3.25034EN0220761034M192
7610.81315.15535EN0220329070M173
777.61590.46056EW0221628397G183
7814.328132.23876EW0236193688G181
7913.033140.36820EW0220807454G159
8010.791151.15127EN1045270925M152
818.723147.34738EN1045473354M164
    EN0244024694M161
823.732142.62527EN1045473354M164
834.711155.90055EW0220505167F198
84−3.90999.50235EN1046573179M213
    EW0221499071G252
852.90073.27644EN0219094286M218
86−7.5525.61140EN1004533697M156
    EN0235553800M183
87−2.610−6.04992EN0220760132M127
    EN0242841854M50
88−3.996−55.94939EW0243855215G242
891.488−75.08779EW0224044850G202
90−2.349−76.26531EN0242044151M214
91−0.067−96.08061EN0211414081M250
    EN0231475794M250
92−3.516−136.80226EN0212284006M101
930.413−177.47229EN0212978685M253
94−13.765−163.43725EN0258339504M99
95−27.948−156.24948EN0247738212M278
96−15.4097.63368EN0220499437M173
97−23.30618.36155EN0255168898M160
    EN0250388819M238
98−12.12698.05427EN0247139401M127
99−22.65596.16225EN0246880606M139
100−19.68592.23825EN0234195338M233
    EN0247139653M136
101−19.70796.85818EN0221499527M107
102−17.84095.46327EN0247053204M130
103−9.766111.41166EN1016176428M103
104−10.014122.00039EN1016031842M173
105−26.521117.897105EN1016090430M114
    EN0264071222M208
    EN0251488753M54
106−16.017151.96146EN1014966566M198
107−14.224171.27543EN1014822402M91
108−10.362164.44038EN0250538312M65
109−33.23369.96723EN1013721837M209
110−39.44553.02753EN1034260978M96
111−41.27143.52741EN1034636136M78
112−30.86243.31025EN0235042831M156
113−31.12023.86865EN1034433492M112
    EN0220023562M256
114−38.34325.606138EN0235212049M171
    EN0235296526M183
115−33.94112.69381EN0235382112M160
    EN0250675342M160
116−38.030−71.31720EN0239249914M171
117−33.997−70.22544EN0238953124M160
    EN0238953122M160
118−37.659−83.11554EN0229280969M114
119−35.238−85.49933EN0229280765M107
    EN0229065549M137
120−30.889−101.18942EN0242040912M199
121−41.112−152.78560EN0262859538M155
    EN0227549830M262
122−45.166−167.94334EN0245441590M111
    EN0219610891M145
123−44.667−165.36328EN0219610891M145
124−49.316−145.12997EN0263232225M198
125−49.013−75.93850EN0239250448M105
126−49.167−73.89027EN0239250452M105
127−57.842−5.53945EN1034691196M154
128−51.67224.56860EN1034576600M135
129−49.74525.98341EN1034461127M143
130−53.18757.48039EN1034173455M130
131−44.16676.88647EN0254534297M175
132−48.40174.28246EN0252326154M124
133−51.627108.98574EN1012943683M123
134−56.575124.19021EN0231656048M127
135−68.303−3.97341EN1017417089M163
136−72.837−9.92921EN1035007027M178
137−79.49556.11640EN0238362031M185
13858.282−177.45291EW0213155056G188
13961.187−62.44784EW1005774359F105
    EW1008396467L312
14062.591−22.17920EW0228025130G114
14161.702−20.62922EW0228025145G115
14252.717−26.25830EW0225863950G157
    EW0251313469G149
14355.4297.57467EW0220330559G181
14454.6649.81253EW0220330559G181
14561.87173.14332EW0254856578G226
14662.57072.83134EW0234537340G257
14748.653179.44039EW0260675925I116
14845.836171.54043EW0260819956G163
14943.798159.17879EW0248490647I175
    EW0220461325G75
    EW0248519506G164
15042.418154.84252EW1045500922G184
15140.460152.95957EW0263756769G174
    EW0220590967G76
    EW0220591013G82
15244.096152.10328EW0220590936G73
    EW0235811292G159
15339.676165.74336EW0235641649G182
    EW0220331770G86
15436.569166.70840EW0260906588G154
    EW0235429532G219
15536.904166.06947EW0260906588G154
    EN1014762926M30
15638.522159.29536EW0220418202G89
15730.417157.20034EW0235769087G233
15835.901145.35040EW0220720658G87
    EW0235938728G204
15943.885143.28060EW0248692273I178
    EW0220720543G73
16038.696140.66530EW0220763820G81
16146.102134.91736EW0248836341I165
16215.78293.80894EW0221541858G152
    EW0246937109G220
16324.76030.29355EW1014300124G224

Note. For additional details on lobate scarp dip directions, see Table A4.

Download table as:  ASCIITypeset images: 1 2 3 4

Table A2. Craters that Exhibit Apparent Straight Rim Segments, Chi-squared Test Results, and Identified PICs

Crater IDChi-squared Test p-values 
 8° Bins8° Bins, Shifted 4°16° Bins16° Bins, Shifted 8°PIC
10.08540.83560.048220.6288Yes
20.1270.59090.016880.346Yes
30.005330.070921.22 × 10−6 0.1927Yes
4<2.20 × 10−16 <2.20 × 10−16 <2.20 × 10−16 <2.20 × 10−16 Yes
100.6580.51390.13130.6288No
110.4030.48930.024370.7751Yes
120.8100.91950.42410.9108No
130.07560.043390.018540.1052Yes
150.9530.78260.018540.9943Yes
160.01550.041881.16 × 10−5 0.1423Yes
172.5 × 10−3 0.036920.0011960.2647Yes
180.06180.014330.035290.5723Yes
210.03390.29170.072280.04627Yes
220.01120.015521.237 × 10−3 0.1714Yes
250.4040.66770.25110.7243No
270.2030.90930.096690.7343No
330.1170.39679.473 × 10−3 0.4472Yes
340.01480.35170.11650.4938Yes
360.5570.19660.53040.308No
370.5830.64530.54610.2336No
380.02540.28646.059 × 10−3 0.4969Yes
390.1190.62470.10530.3678No
400.1210.90150.10160.3733No
440.9560.50430.33430.9395No
480.3360.19920.0950.2522No
490.2370.36850.017910.5782Yes
500.6220.76190.53730.8716No
510.02470.046586.448 × 10−3 0.03071Yes
520.3610.29960.24490.2994No
550.2820.2820.38550.4756No
590.1830.36430.29330.6253No
610.4570.53250.47080.4171No
640.4420.49470.11490.7435No
670.4580.11250.41960.5119No
680.2640.78980.50930.3422No
690.8870.65140.49380.8707No
700.2090.60920.067190.4907No
740.5620.66970.12320.9439No
760.1760.42080.019780.2011Yes
800.5740.5740.67130.8193No
810.5420.35460.40890.4907No
820.39050.7440.089950.9473No
840.1110.77950.042380.4803Yes
860.8060.74280.22680.9272No
940.6100.6100.48540.3857No
950.2840.74620.098760.1855No
980.5690.41440.36360.9311No
990.3490.95440.19730.5132No
1000.5210.86220.73650.5321No
1010.8580.54140.82770.9123No
1020.06520.021550.040150.03779Yes
1080.5250.65220.20820.3593No
1090.8710.58860.46590.9149No
1150.04080.29870.061090.2398Yes
1164.96 × 10−4 0.021390.00017877.27 × 10−3 Yes
1220.2810.34720.027250.09631Yes
1270.3050.85130.32730.9563No
1340.5250.69870.041320.9377Yes
1350.4380.64420.54850.6288No
1402.87 × 10−3 0.10110.0048020.06413Yes
1410.1190.69830.097810.7406No
1452.31 × 10−3 0.40410.05490.4275Yes
1480.1290.81249.378 × 10−3 0.8524Yes
1520.1610.074385.24 × 10−3 0.3582Yes
1530.9380.96440.330.9755No
1560.4740.5880.15030.8912No
1570.8540.52380.93120.8438No
1580.5400.26440.20160.9275No
1590.03840.31631.698 × 10−4 0.274Yes

Download table as:  ASCIITypeset image

Annotated portions of MDIS images covering each PIC identified are shown in Figure A1 for craters 1–22, Figure A2 for craters 33–134, and Figure A3 for craters 140–159. Lobate scarps are visible in some of these images; however, in many cases, the scarps are most visible some distance from each crater due to overprinting of the scarps by impact ejecta. See Figure 5 for context images (with lower resolutions) that show PICs with their adjacent lobate scarps in single MDIS images.

Figure A1.

Figure A1. Identified PICs for craters 1–22 (see Table A2). Cyan arrows indicate examples of straight rim segments. See Table A1 for information on PIC locations, diameters, and MDIS image coverage. See Figures A2 and A3 for additional identified PICs.

Standard image High-resolution image
Figure A2.

Figure A2. Identified PICs for craters 33–134 (see Table A2). Cyan arrows indicate examples of straight rim segments. See Table A1 for information on PIC locations, diameters, and MDIS image coverage. See Figures A1 and A3 for additional identified PICs.

Standard image High-resolution image
Figure A3.

Figure A3. Identified PICs for craters 140–152 (see Table A2). Cyan arrows indicate examples of straight rim segments. See Table A1 for information on PIC locations, diameters, and MDIS image coverage. See Figures A1 and A2 for additional identified PICs.

Standard image High-resolution image

Annotated rose diagrams for all PICs analyzed and their associated lobate scarps are provided in Figure A4 for craters 1–11, Figure A5 for craters 13–18, Figure A6 for craters 21–38, Figure A7 for craters 49–102, Figure A8 for craters 111–140, and Figure A9 for craters 145–159. See Table 1 for a summary of PIC straight rim segment azimuths and azimuths of adjacent lobate scarps.

Figure A4.

Figure A4. Rose diagrams showing the azimuth distributions of PIC rims and adjacent lobate scarps for craters 1–11 (Table A1). Modes are represented by blue petals, and red stars denote similar orientations between PIC straight rim segments and lobate scarps.

Standard image High-resolution image
Figure A5.

Figure A5. Rose diagrams showing the azimuth distributions of PIC rims and adjacent lobate scarps for craters 13–18 (Table A1). Modes are represented by blue petals, and red stars denote similar orientations between PIC straight rim segments and lobate scarps.

Standard image High-resolution image
Figure A6.

Figure A6. Rose diagrams showing the azimuth distributions of PIC rims and adjacent lobate scarps for craters 21–38 (Table A1). Modes are represented by blue petals, and red stars denote similar orientations between PIC straight rim segments and lobate scarps.

Standard image High-resolution image
Figure A7.

Figure A7. Rose diagrams showing the azimuth distributions of PIC rims and adjacent lobate scarps for craters 49–102 (Table A1). Modes are represented by blue petals, and red stars denote similar orientations between PIC straight rim segments and lobate scarps.

Standard image High-resolution image
Figure A8.

Figure A8. Rose diagrams showing the azimuth distributions of PIC rims and adjacent lobate scarps for craters 115-140 (Table A1). Modes are represented by blue petals, and red stars denote similar orientations between PIC straight rim segments and lobate scarps.

Standard image High-resolution image
Figure A9.

Figure A9. Rose diagrams showing the azimuth distributions of PIC rims and adjacent lobate scarps for craters 145–159 (Table A1). Modes are represented by blue petals, and red stars denote similar orientations between PIC straight rim segments and lobate scarps.

Standard image High-resolution image

Table A3. Descriptions of Analyzed Lobate Scarp Proximities to Craters

Crater IDDistance from Measured Lobate Scarp Section to Crater Rim (km)Intersection between Crater and Lobate Scarp Fault?Crater Diameter (km)
 Scarp AScarp BScarp AScarp B 
PICs
11Yes, edge21
21515Yes, edgeNo22
30Yes, center22
441Yes, edgeYes, center23
114Yes, edge40
139Yes, edge32
1511No20
164167Yes, edgeNo67
1735Yes, edge67
1814Yes, center41
215446NoNo47
2233No26
3325Yes, center24
3412Yes, edge74
3838No38
491112Yes, centerYes, center35
515Yes, center46
7618Yes, center35
848Yes, edge35
102133NoNo27
11556Yes, edge81
116213Yes, centerNo20
122917Yes, edgeYes, edge34
1343Yes, center21
14039Yes, edgeYes, edge20
1458Yes, center32
1481239Yes, centerYes, center43
15218Yes, center28
15921No60
 
CICs
 
52416NoYes, center41
61810Yes, centerYes, center26
77No25
835Yes, center27
913No31
1016Yes, center30
123No20
1415Yes, center34
194Yes, center32
2076Yes, center45
2320Yes, center80
2440Yes, center45
2560Yes, center40
2636Yes, center77
2726No45
2813Yes, center41
298Yes, edge27
3041No48
3126No35
329Yes, edge24
3512Yes, center43
360Yes, center26
3719No27
3911No27
4012No24
4119No28
429Yes, center25
4317Yes, center26
4410Yes, edge28
4517No45
468Yes, center39
4717No54
481Yes, center35
5016Yes, center30
523No30
53144Yes, edgeYes, edge24
542Yes, edge24
5510Yes, center25
5625Yes, center76
571423Yes, centerNo42
583011Yes, centerYes, center59
593Yes, center28
6019No78
61817Yes, centerNo24
623No26
6316Yes, center30
6417Yes, center30
6530No33
6620Yes, center33
671810Yes, centerNo24
684037NoNo47
69513Yes, edgeYes, center35
70810Yes, edgeYes, center45
7111Yes, center27
7213Yes, center34
73151Yes, edgeYes, center25
7418Yes, center35
7581Yes, centerYes, center34
7722Yes, center56
7811Yes, center76
7934NoNo20
801315NoYes, center27
818Yes, edge38
82622Yes, centerNo27
8330Yes, center55
855Yes, center44
8635No40
8773Yes, center92
884Yes, center39
8916Yes, center79
908Yes, edge31
9114Yes, center61
921438Yes, centerNo26
931823NoNo29
941015NoYes, center25
9532No48
9637No68
9752Yes, center55
98142NoYes, center27
9913Yes, center25
10021Yes, center25
10196NoNo18
103128Yes, centerYes, center66
1041Yes, edge39
1058Yes, center105
106229NoNo46
107810Yes, centerYes, edge43
1087Yes, center38
1095No23
1101333NoNo53
11126Yes, center41
11258Yes, centerYes, center25
1133Yes, center65
11432Yes, center138
11736No44
118197NoYes, center54
11926Yes, center33
12010Yes, center42
1211642Yes, centerNo60
12313Yes, center28
12454Yes, center97
1256Yes, center50
12617Yes, center27
1274No45
128166Yes, centerYes, center60
12939No41
13016Yes, center39
1312025Yes, centerYes, center47
1320Yes, center46
1335Yes, center74
1357No41
13615No21
1377No40
13854No91
139856Yes, centerYes, center84
14117Yes, centerYes, center22
14229No30
14364No67
14420Yes, center53
1462418NoYes, edge34
1474Yes, center39
1493742NoNo79
1501538Yes, centerYes, center52
1512842Yes, centerYes, center57
1531113Yes, centerYes, center36
15490Yes, center40
1558No47
15613No36
1572332NoNo34
1581320Yes, centerYes, edge40
16014No30
16114Yes, center36
16297No94
16317No55

Download table as:  ASCIITypeset images: 1 2 3

Table A4. Dip Directions of Each Lobate Scarp Analyzed

Crater IDLobate Scarp Dip Direction
 Scarp AScarp B
1SE
2NWSE
3SW
4ESE
5NWNW
6WSW
7S
8SE
9NW
10NE
11W
12NE
13E
14SE
15SE
16EE
17NW
18SE
19NW
20NW
21ESE
22SW
23SE
24E
25SW
26E
27NW
28W
29NW
30NW
31SE
32SW
33E
34SE
35SE
36SE
37SW
38SW
39NE
40NW
41W
42E
43SE
44W
45W
46SE
47NE
48SE
49WSW
50W
51NE
52E
53NWE
54SE
55E
56SW
57EE
58WNW
59NW
60SE
61ENW
62SE
63SE
64SW
65NW
66SE
67NENW
68NWSW
69NWE
70NENE
71SE
72NW
73SESW
74E
75NWNW
76SE
77W
78SE
79SWW
80WSW
81NW
82ESE
83NW
84ESE
85E
86NE
87SW
88SW
89E
90E
91NW
92NENW
93NESE
94ENE
95SE
96NE
97NE
98NENE
99SE
100SE
101SEE
102NESE
103ENE
104W
105W
106WSW
107ENE
108E
109NW
110NWNW
111NW
112ENE
113SE
114NW
115NW
116WSE
117NE
118NWNE
119NW
120NE
121NENW
122NWSW
123W
124NE
125NE
126NW
127SW
128SEN
129NW
130SE
131NWNE
132SE
133NE
134NE
135SW
136NE
137E
138W
139SWNE
140SWNW
141SWNW
142W
143NE
144NW
145SW
146NESE
147SE
148SESE
149NWSW
150SWNW
151NENW
152SE
153ENE
154NE
155NE
156W
157WNW
158NESE
159SE
160SE
161NE
162NW
163NE

Note. See Table A1 for locations of each crater and adjacent lobate scarp.

Download table as:  ASCIITypeset images: 1 2 3 4

Footnotes

Please wait… references are loading.
10.3847/PSJ/ad1fff