Open Access
How to translate text using browser tools
18 May 2018 Ichnotaxonomy of the Cambrian Spence Shale Member of the Langston Formation, Wellsville Mountains, Northern Utah, Usa
Sean R. Hammersburg, Stephen T. Hasiotis, Richard A. Robison
Author Affiliations +
Abstract

The Spence Shale of northern Utah is the oldest North American middle Cambrian (∼506–505 Ma) Burgess Shale-type (BST) deposit and, unlike previously thought for BST deposits, has a very diverse ichnofauna. Twenty-four ichnogenera and 35 ichnospecies were identified: Archaeonassa (A. fossulata and A. jamisoni isp. nov.), Arenicolites carbonaria, Aulichnites, Bergaueria (B. hemispherica and B. aff. perata), Conichnus conicus, Cruziana (C. barbata and C. problematica), Dimorphichnus, Diplichnites (D. cf. binatus, D. gouldi, and D. cf. govenderi), Gordia marnia, Gyrophyllites kwassizensis, Halopoa aff. imbricata, Lockeia siliquaria, Monomorphichnus (M. bilinearis, M. lineatus, and M. cf. multilineatus), Nereites cf. macleayi, Phycodes curvipalmatum, Phycosiphon incertum, Planolites (P. annularius, P. beverleyensis, and P. montanus), Protovirgularia (P. dichotoma and P. cf. pennatus), Rusophycus (R. carbonarius, R. cf. pudicus, and R. cf. cerecedensis), Sagittichnus lincki, Scolicia, Taenidium cf. satanassi, Teichichnus cf. nodosus, and Treptichnus (T. bifurcus, T. pedum, and T. vagans). The ichnofossils comprise three ichnocoenoses—Rusophycus-Cruziana, Sagittichnus, and Arenicolites-Conichnus—representing dwelling, deposit- and filter-feeding, grazing, locomotion, and predation behaviors of organisms (e.g., annelid worms and trilobites). Two ichnofossil associations are suggestive of predation: (1) Planolites terminating at a Rusophycus; and (2) Archaeonassa crosscutting a Taenidium. The Spence Shale ichnofauna represent a distal Cruziana Ichnofacies and depauperate, distal Skolithos Ichnofacies. A new ichnospecies of Archaeonassa is proposed, A. jamisoni isp. nov., and Ptychoplasma (Protovirgularia) vagans is herein transferred to Treptichnus. This study is the first ichnotaxonomic study of the Spence Shale and North American BST deposits and shows highly diverse ichnofaunas can be present in BST deposits.

INTRODUCTION

Although rare in the fossil record, soft-tissue preservation has provided paleontologists with a detailed glimpse into unique paleoenvironments with even more unique and sometimes bizarre faunas not seen elsewhere. Soft tissues are most commonly preserved as kerogenized carbonaceous films, known as Burgess Shale-type (BST) preservation (e.g., Gaines, Kennedy, & Droser, 2005). A fossilization mode still not well understood, numerous studies of BST deposits (e.g., Butterfield, 1990, 1995; Allison & Brett, 1995; Petrovich, 2001) have tried to delineate and understand the mechanics and paleoenvironmental conditions necessary for BST production, including whether or not the absence of ichnofossils is necessary. Understanding the physicochemical controls can help refine depositional, paleoenvironmental, and paleoecological interpretations of BST deposits. Ichnofossils, however, can be used as proxies for paleoenvironmental and physicochemical conditions (e.g., sedimentation rate, benthic paleooxygenation, nutrients, depositional energy, etc.) present during and after deposition, even when body fossils are absent (e.g., Bromley, 1996; Hasiotis & Platt, 2012), and, therefore, can aid in understanding BST production.

Rare throughout the middle and upper Proterozoic and lower Phanerozoic, most BST deposits occur globally in lower and middle Cambrian (Terreneuvian—Series 3) rocks with most middle Cambrian BST deposits confined to North America (Conway Morris, 1992; Butterfield, 1995; Garson & others, 2012). The most well-known North American BST deposits include (by age), the Spence Shale (∼506–505 Ma), Burgess Shale, Wheeler Formation, and Marjum Formation (Gaines & Droser, 2005; Garson & others, 2012). The Spence Shale of northern Utah (Fig. 1) is the oldest among North American BST deposits with unique preservation of both soft tissues and numerous ichnofossils sometimes in the same stratigraphic intervals or in direct contact (Garson & others, 2012).

Figure 1.

Map of study area and collection sites from the Spence Shale. 1, Spence Shale collection localities in northern Utah and southern Idaho, shaded area denotes presence of the Spence Shale (modified from Liddell, Wright, & Brett, 1997); 2, Topographic map of the Wellsville Mountain area, north of Brigham City with Langston Formation outcrops shaded (modified from Jensen & King, 1999).

f01_01.jpg

The purpose of this study is to: (1) document the ichnofossils and ichnodiversity of the Spence Shale; (2) establish ichnocoenoses and assign ichnofacies; and (3) compare the Spence Shale ichnofauna to ichnofaunas present in other BST and Cambrian-aged deposits. Detailed ichnotaxonomic studies on BST deposits are necessary so that ichnocoenoses and ichnofacies can be established to further interpret the physicochemical controls that determined faunal types and the type and degree of bioturbation.

This is the first study to conduct a detailed ichnotaxonomic examination of ichnofossils in a North American BST deposit, which will help form a baseline for BST deposits. No significant ichnotaxonomic work exists and only a few reports of ichnofossils are available for the Wheeler and Marjum formations (Ubaghs & Robison, 1985; Robison, 1991; Gaines & Droser, 2005; Gaines, Kennedy, & Droser, 2005). Similarly, very little ichnotaxonomic work exists from the Burgess Shale (e.g., Caron & others, 2010; Mángano, 2011; Minter, Mángano, & Caron, 2012). There are several ichnotaxonomic studies from lower and middle Cambrian BST deposits of China: the Chengjiang Formation (e.g., Zhang & others, 2007; Huang & others, 2014) and the Kaili Formation (e.g., Yang, 1994; Yang & Zhao, 1999; Wang & others, 2004, 2009; Lin & others, 2010). Ichnotaxonomic comparisons between Chinese and North American BST deposits, as well as others, help establish the range of paleoenvironments and physicochemical conditions in which BST fossils had been produced.

BACKGROUND

The Spence Shale was first described by Walcott (1908) from the Spence Gulch, southeastern Idaho, after a Bear River Range resident, R.S. Spence, began a 10-year correspondence in 1896 sending numerous well-preserved fossils to Walcott (Resser, 1939). Described as a 30-foot-thick (9.1 m) “argillaceous shale with sandy shale,” the Spence Shale was interpreted as the basal member of the Ute Formation of Idaho (Walcott, 1908; Resser, 1939). Maxey (1958) later placed the Spence Shale as the middle member of the Langston Formation between the Naomi Peak Limestone (basal) and High Creek Limestone (upper) members. Oriel and Armstrong (1971), however, placed the Spence Shale as a tongue deposit within the Lead Bell Shale of Idaho. Subsequent authors have followed Maxey (1958) for units outcropping within Utah (e.g., Hintze & Robison, 1975; Robison, 1976; Conway Morris & Robison, 1988; Liddell, Wright, & Brett, 1997; Garson & others, 2012); whereas, Oriel and Armstrong (1971) has remained in use for outcrops in Idaho (e.g., Palmer & Campbell, 1976; Liddell, Wright, & Brett, 1997). Robison (1991) proposed that the Spence Shale be elevated to formation rank, but to date, no author has accepted this proposal (Liddell, Wright, & Brett, 1997; Garson & others, 2012).

Middle and upper Cambrian units of the Great Basin of Utah were deposited in a north—south-trending (present-day orientation) carbonate belt, flanked by inner (eastern) and outer (western) detrital belts (Palmer, 1960; Robison, 1960) (Fig. 2). The Spence Shale was deposited mostly within the outer detrital belt and some of the middle carbonate belt (Robison, 1960; Liddell, Wright, & Brett, 1997; Garson & others, 2012). Palmer and Campbell (1976) proposed three biofacies for the Langston Formation and equivalent strata: (1) low-diversity, restrictedshelf biofacies corresponding to deposition in the inner detrital belt; (2) high-diversity, platform-margin to open-shelf biofacies corresponding to deposition in the middle carbonate belt; and (3) deep-shelf or basinal, low-diversity biofacies characterized by agnostoid and oryctocephalid trilobites. Robison (1976) showed that the agnostoid and polymeroid trilobite distributions of the Langston Formation correlated with the carbonate and detrital belts similar to the Palmer and Campbell (1976) biofacies. The restricted-shelf biofacies includes the sandy units of the Naomi Peak Limestone Member (also known as Twin Knobs Formation of Idaho), whereas, the platform-margin to open-shelf biofacies corresponds to most of the limestones and shales of the Langston Formation, and the deep-shelf biofacies corresponds to the shales at the Oneida Narrows locality (Liddell, Wright, & Brett, 1997).

Several models have been proposed for the production of BST and each suggests a dominant environmental physicochemical factor(s): (1) rapid burial and benthic anoxia (Conway Morris, 1986); (2) clay-rich sediment to allow adsorption of enzymes into surrounding clays and inhibit decomposition (Butterfield, 1990, 1995); (3) oscillations between benthic anoxia and dysoxia (Allison & Brett, 1995); or (4) iron mineral-rich sediment to allow iron (II) adsorption and inhibit bacterial decomposition (Petrovich, 2001). Gaines and Droser (2005) and Gaines, Kennedy, and Droser (2005) developed a new model for BST from the Wheeler Formation of central Utah requiring siliciclastic clay-dominant, mixed siliciclastic-carbonate sediment with low original porosity, proximity to both oxic and anoxic bottom waters, and little to no bioturbation. Gaines and Droser (2010) used ichnofabric indices to confirm the Gaines, Kennedy, and Droser (2005) model for the Wheeler and Marjum formations and found that benthic anoxia was necessary for BST production. Similarly, Garson and others (2012) used bioturbation patterns via ichnofabric indices to interpret the Spence Shale benthic paleooxygenation and found that significant bottom-water oxygenation occurred and was persistent for some periods and rapidly alternated between anoxic and oxic conditions during others.

GEOLOGIC SETTING

During the middle Cambrian, present-day northern Utah was located on the northwestern margin of Laurentia (Fig. 3; Liddell, Wright, & Brett, 1997). The Spence Shale is the early middle Cambrian (Series 3, Stage 5), middle member of the Langston Formation in northern Utah stratigraphy (Maxey, 1958; Liddell, Wright, & Brett, 1997; Garson & others, 2012; Peng, Babcock, & Cooper, 2012). In Utah, the Langston Formation is underlain by the Geertsen Canyon Quartzite of the Neoproterozoic—lower Cambrian Brigham Group and overlain by the Ute Formation (Fig. 4; Maxey, 1958; Liddell, Wright, & Brett, 1997). In the Wellsville Mountain area, the Spence Shale is underlain by the Naomi Peak Limestone Member (Twin Knobs Formation of Idaho) and overlain by the High Creek Limestone Member (Maxey, 1958; Liddell, Wright, & Brett, 1997).

The Spence Shale is a 50—65-m-thick, gray to black, calcareous shale interbedded with peloidal—oolitic limestone intervals and sandy stringers (Fig. 5) deposited on a ramp setting, shifting from proximal to distal as time progressed (Liddell, Wright, & Brett, 1997; Garson & others, 2012) (see Fig. 2). The Spence Shale contains several stacked, shallowing parasequences that lead to deposition of peloidal, oolitic, and nodular limestone intervals (Liddell, Wright, & Brett, 1997).

The Spence Shale has an abundant and diverse hard-bodied fauna, including, agnostoid and polymeroid trilobites, articulate and inarticulate brachiopods, eocrinoids, mollusks, and sponges (Walcott, 1908; Resser, 1939; Gunther & Gunther, 1981; Babcock & Robison, 1988; Robison, 1991; Liddell, Wright, & Brett, 1997; Sprinkle & Collins, 2006; Briggs & others, 2008). It also contains a diverse soft-bodied fauna, including, algae, annelids, and soft-shelled arthropods (Robison, 1969, 1991; Briggs & Robison, 1984; Conway Morris & Robison, 1988; Liddell, Wright, & Brett, 1997). Traditionally, ichnofossils and BST fossils are not thought to normally occur in close proximity to each other but to be deposited in exclusive zones of oxia—dysoxia and anoxia, respectively (Allison & Brett, 1995). There are, however, increasing reports of ichnofossils and BST fossils occurring together, recording interactions between tracemaking organisms and BST fossils, suggesting more dynamic paleoenvironmental conditions during deposition (e.g., Zhang & others, 2007; Wang & others, 2009). Prior to this study, only a few ichnofossils were reported or described from the Spence Shale, including Brooksella, coprolites, Cruziana, Gyrophyllites, Neonereites, Palaeophycus, Planolites, Rusophycus, Tasmanadia, and Treptichnus (Robison, 1969, 1991; Willoughby & Robison, 1979; Ubaghs & Robison, 1985; Conway Morris & Robinson, 1986).

Figure 2.

Depositional and biofacies models, ichnocoenoses, ichnofacies, and physicochemical controls of the Langston Formation and equivalent units (modified from Palmer & Campbell, 1976; Robison, 1976; Liddell, Wright, & Brett, 1997).

f02_01.jpg

Figure 3.

Middle Cambrian paleogeography of Laurentia with Spence Shale location (star) (modified and redrawn with permission from Colorado Plateau Geosystems © 2007).

f03_01.jpg

Figure 4.

Cambrian geologic time scale and correlated biostratigraphy of Series 3 Cambrian Burgess Shale-type units of China and North America. 1, Cambrian Period geologic time scale with series and stages/ages (redrawn from Peng, Babcock, & Cooper, 2012); 2, Cambrian Series 2–3 trilobite biozonation correlation between South China and North America (i.e., Laurentia) with Burgess Shale-type deposits; BC, CA = British Columbia, Canada (modified from Liddell, Wright, & Brett, 1997; Collom, Johnston, & Powell, 2009; Lin & others, 2010; Robison & Babcock, 2011; Peng, Babcock, & Cooper, 2012); 3–4, Northern Utah and southern Idaho stratigraphic correlation (modified from Liddell, Wright, & Brett, 1997).

f04_01.jpg

ABBREVIATIONS

The abbreviations used in this study include; KUMIP, University of Kansas Museum of Invertebrate Paleontology; IBGS, IchnoBioGeoScience Research Group (University of Kansas). Key to fossil collection naming; YY-A-XXX [YY: collector and donor (LG; Lloyd Gunther, PJ; Paul Jamison); A: depositional realm (C: continental, M; marine); XXX: three-digit specimen number].

MATERIALS AND METHODS

Material for this study (Fig. 624) comes from Spence Shale outcrops in the Wellsville Mountains of northern Utah, USA. Specimens were collected and donated to the KUMIP and IBGS collections by Lloyd and Val Gunther, Paul Jamison, Phillip Reese, and Richard A. Robison. Specimens were measured using nondigital Vernier calipers (0.1 mm accuracy). Long or winding traces were measured by a waxed string, which was then measured with calipers. ImageJ (v. 1.48; USNIH, 2015) analysis software was used to measure V-shaped angles of striation patterns, dimensions of smaller specimens, and grain sizes. Several specimen slabs were prepared for in-laboratory examination and photography with a 2.0% HCl acid solution to dissolve thin surficial carbonate deposits obscuring underlying traces. Unpolished sections of cut samples were wetted with glycerin and photographed. Specimens were examined in hand sample and using a Nikon SMZ1000 binocular light microscope. Specimen photographs were taken with a mounted Sony Cyber-shot DSC-HX200V camera or a Nikon DXM1200 digital camera attached to the Nikon SMZ1000 microscope. Photographs were processed with Adobe Photoshop™ Creative Cloud (CC) version.

Figure 5.

Stratigraphy of the Spence Shale with ichnofossil placement (modified from Liddell, Wright, & Brett, 1997).

f05_01.jpg

Ichnological assessments were made following several methodologies. Descriptions of architectural and surficial morphology follow Hasiotis and Mitchell (1993), Bromley (1996), and Hasiotis (2004, 2008). Trackways were described using the terminology of Trewin (1994), Keighley and Pickerill (1998), and Minter, Braddy, and Davis (2007). Samples with visible bedding or laminations and bioturbation were analyzed via the Ichnofabric Index (ii; Droser & Bottjer, 1986). Bedding planes were analyzed with the Bedding-Plane Bioturbation Index (BPBI); Miller & Small, 1997). Since this material was collected and donated by private collectors, the establishment of truly representative ichnocoenoses is difficult. Each examined slab specimen is itself a unique ichnocoenosis and represents a single community of traces. Overarching ichnocoenoses were constructed via reoccurring ichnofossil associations following Pemberton and others (2001) and Jackson, Hasiotis, and Flaig (2016).

SYSTEMATIC ICHNOLOGY

Ichnogenus ARCHAEONASSA Fenton & Fenton, 1937a

  • Type ichnospecies.—Archaeonassa fossulata Fenton & Fenton, 1937a.

  • Emended Diagnosis.—Short, round- to ovoid-shaped or elongated trails or burrows commonly deeper at one end, and may grade into indistinct V-shaped trails; concave to slightly convex furrow flanked by pair of convex ridges, central furrow typically wider than ridges; lateral convex ridges may be smooth or ornamented with oblique to transverse striations or smaller lobes (Fenton & Fenton, 1937a; Buckman, 1994).

  • Discussion.—Fenton and Fenton (1937a) established Archaeonassa for elongate, concave furrows with flanking convex ridges produced by snails and other gastropods from the lower Cambrian Mount Whyte Formation of British Columbia. Häntzschel (1975) placed Archaeonassa in the Scolicia Group, but was not placed in synonymy with Scolicia because Archaeonassa lacks any complex backfill diagnostic to Scolicia (Buckman, 1994). Buckman (1994) reviewed Archaeonassa and considered it the senior synonym of Scolicia vada Chamberlain, 1971, and some specimens of Palaeobullia Götzinger & Becker, 1932. Yochelson and Fedonkin (1997) rejected this synonymy, however, partly because Buckman (1994) did not include the original type material of Archaeonassa while also including ornamented lateral ridges.

  • Archaeonassa is generally interpreted as a gastropod locomotion or grazing trace ((Fenton & Fenton, 1937a; Buckman, 1994; Jensen, Droser, & Gehling, 2005). Yochelson and Fedonkin (1997), however, suggested that Archaeonassa was not produced by mollusks but did not suggest any other producers. Trilobites and echinoids have also been suggested as possible tracemakers (Buckman, 1994). Jensen, Droser, and Gehling (2005) pointed out that such protists as foraminifera can make traces similar to Archaeonassa but are rarely considered as producers. Buchanan and Hedley (1960, p. 557–558) did not figure any ichnofossils (i.e., only provided drawings of the pseudopodial systems used by forams), but provided a description of foram-produced furrows: “… a furrow is left in the sand as a result of the leading edge of the test being preceded by a raised mound or ‘bow-wave’ of sand.” This description, however, does match most Archaeonassa descriptions. Archaeonassa has mostly been reported from shallow-marine deposits (e.g., tidal flats), as well as from continental deposits (e.g., delta front, fluvial, and lacustrine; e.g., (Fenton & Fenton, 1937a; Buatois & Mángano, 2002, 2007; Mángano, Buatois, & Muñiz Guinea, 2005). Archaeonassa was recently reported from flysch deposits from India (Khaidem, Rajkumar, & Soibam, 2015); however, those specimens are overlapping, bilobate, convex epireliefs, likely Crossopodia M‘Coy, 1851 or Gyrochorte Heer 1865 in Heer 1864–1865. Archaeonassa ranges from the Ediacaran to recent (e.g., (Fenton & Fenton, 1937a, Jensen, Droser, & Gehling, 2005; Buckman, 1994; Martin, 2013).

  • ARCHAEONASSA FOSSULATA ((Fenton & Fenton, 1937a)
    Figure 6.1

  • Material.—IBGS PJ-M-027: one specimen, Miners Hollow; IBGS PJ-M-033: one specimen (part and counterpart), Miner's Hollow.

  • Diagnosis.—Concave to slightly convex furrow flanked by pair of convex ridges; central furrow wider than flanking ridges; the lateral convex ridges may be smooth or ornamented with oblique to transverse striations or smaller lobes ((Fenton & Fenton, 1937a; Buckman, 1994).

  • Description.—Convex furrow with concave lateral ridges in hyporelief (IBGS PJ-M-027) and concave furrow with convex lateral ridges in epirelief (IBGS PJ-M-033). Furrows 37.4-40.2 mm long, 3.7–5.4 mm wide, and 1.4 mm deep; lateral ridges 0.6–2.8 mm wide.

  • Occurrence.—Gray to slightly blue-gray (weathered to tan), calcareous and micaceous silty shale.

  • Associated ichnotaxa.Gyrophyllites kwassizensis, Nereites cf. macleayi, Planolites montanus, and Treptichnus pedum.

  • Discussion.—Specimens were assigned to Archaeonassa fossulata based on the simple and smooth furrows flanked by lateral ridges in epirelief (Fig. 6.1). The specimen of Archaeonassa on IBGS PJ-M-027 occurs as a convex furrow with concave lateral ridges in hyporelief. The width and depth of the furrow and lateral ridges are not uniform. The furrow and ridges are narrower and shallower on one end than on the other, suggesting the tracemaker may have been burrowing obliquely through the sediment. The specimen on IBGS PJ-M-033 occurs as a concave furrow with convex lateral ridges (in part and counterpart).

  • Figure 6.

    Archaeonassa, Arenicolites, Aulichnites, and Bergaueria specimens from the Spence Shale. 1, Archaeonassa fossulata with convex ridges (black arrows) and concave furrow (white arrow) in concave epirelief, IBGS PJ-M-033; 2, Archaeonassa jamisoni isp. nov. with holotype (arrow), in convex and concave epirelief, IBGS PJ-M-005, Miner's Hollow float; 3–4, Arenicolites carbonaria, IBGS PJ-M-003, Cataract Canyon; 3, Arenicolites carbonaria apertures in concave epirelief; 4, Arenicolites carbonaria in full relief; 5, Concave hyporelief of Aulichnites isp. (black arrow) terminating at Lockeia siliquaria (white arrow) and Protovirgularia cf. pennatus (left center) in convex hyporelief, IBGS PJ-M-019, Miner's Hollow; 6, Bergaueria hemispherica near the termination of Teichichnus c.f. nodosus (arrow) in convex hyporelief, IBGS PJ-M-025, Cataract Canyon; scales in cm.

    f06_01.jpg

    ARCHAEONASSA JAMISONI new ichnospecies
    Figure 6.2, Figure 19.3–19.4

  • Material.—IBGS PJ-M-002: two specimens, Miner's Hollow; IBGS PJ-M-005: three specimens, Spence Shale float, Cataract Canyon.

  • Diagnosis.—Smooth, curved, asymmetrical furrow in concave epirelief with or without paired convex lateral ridges along sides of furrow that may merge at furrow terminations with massive fill; infill and lateral ridges may be absent, forming depressions with terracing on trace wall.

  • Description.—Curved, asymmetrical furrow with convex lateral ridges in epirelief. Furrows 47.2–52.6 mm long, 11.2–38.2 mm wide, and 1.4–5.7 mm deep; lateral ridges 2.7–5.2 mm wide and 2.1–3.8 mm thick, merged at furrow terminations.

  • Etymology.—After Paul Jamison, who collected and donated a large number of fossil specimens used in this study.

  • Types.—Holotype: IBGS PJ-M-005; Paratype: IBGS PJ-M-002.

  • Type stratum.—Cambrian, Series 3, Spence Shale Member of the Langston Formation.

  • Type locality.—Miner's Hollow, west side of Wellsville Mountain: T10N, R2W, Sec. 14, NE1/4 SW1/4 and NW1/4 SE1/4 (41° 36′ 4.8″N, 112° 2′ 12.5″W).

  • Repository.—Division of Invertebrate Paleontology, Museum of Natural History and Biodiversity Research Center, University of Kansas, Lawrence, Kansas, USA.

  • Occurrence.—Two lithologies: (1) Tan to light brown, siliciclastic silty shale; and (2) gray, calcareous shale.

  • Associated ichnotaxa.Phycodes curvipalmatum and Taenidium cf. satanassi.

  • Discussion.—Until now, Archaeonassa was monotypic. The closest published morphotype resembling A. jamisoni was from a neoichnological experiment by Jensen, Droser, and Gehling (2005, fig. 2C) that produced asymmetrical undertraces via locomotion of a marine gastropod, Nassarius (Hinia) reticulata (Linnaeus, 1758). They compared the experimental traces to specimens of Archaeonassa from the Ediacaran Ust' Pinega Formation of northwest Russia and the Ediacara Member of the Rawnsley Quartzite, Flinder's Ranges, South Australia. Jensen, Droser, and Gehling (2005) suggested that the Neoproterozoic Archaeonassa represented movement over sandy media and were analogous to the specimens generated by creeping gastropods during their experiment. Sören Jensen (personal communication, 2014) suggested that the A. jamisoni specimens were likely produced by a similar behavior. Martin (2013, fig. 6, p. 266) illustrated a modern moon-snail trace that consisted of a short, concave, asymmetrical furrow flanked by lateral ridges and greatly resembled A. jamisoni (Fig. 6.2). Jean-Bernard Caron (personal communication, 2016) suggested, however, that an ichnofossil interpretation of A. jamisoni is highly dubious due to the wide range of morphology between specimens and may actually be nodular concretions. We disagree with the nodular-concretion interpretation due to the presence of several shallow furrows that widen and deepen proximal to the specimens and taper and shallow out away from them (see Fig. 19.3). We interpret the shallow furrows to be short, entry furrows of a biogenic affinity.

  • Yochelson and Fedonkin (1997) noted that the original description of Archaeonassa contained two morphologies (elongate ribbon traces and rimmed pits) but restricted Archaeonassa to elongate ribbon traces and did not discuss the rimmed pits (resting traces) mentioned by Fenton and Fenton (1937a, p. 454). The rimmed pits of Archaeonassa jamisoni differ from A. fossulata due to the lack of elongate, ribbonlike furrow morphology typical of the ichnospecies. The Spence Shale material presented herein is, therefore, assigned to Archaeonassa based on comparisons to material described in Fenton and Fenton (1937a) and Buckman (1994) and discussions with S. Jensen. We interpret A. jamisoni to represent a combined locomotion and resting trace and possibly even a hunting trace of a gastropod.

  • Ichnogenus ARENICOLITES Salter, 1857

  • Type ichnospecies.—Arenicola carbonaria Binny, 1852 by subsequent designation (Richter, 1924, p. 137).

  • Diagnosis.—Vertical, U-shaped burrows without spreiten, and visible as paired openings in plan view (Fürsich, 1974a; Fillion & Pickerill, 1990).

  • Discussion.—Arenicolites is a U-shaped burrow similar to Diplocraterion Torell, 1870, but lacks spreite between tubes (Hakes, 1976; Fillion & Pickerill, 1990). Ten ichnospecies of Arenicolites are recognized: A. brevis Matthew, 1890; A. carbonaria Binney, 1852; A. compressus (Sowerby, 1829); A. curvatus Goldring, 1962; A. longistriatus Rindsberg & Kopaska-Merkel, 2005; A. naraensis Badve & Ghare, 1978; A. sparsus Salter, 1857; A. statheri Bather, 1925; A. subcompressus (Eichwald, 1860); and A. variabilis Fürsich, 1974a. Arenicolites compressus, A. curvatus, and A. subcompressus have elliptical cross sections, and A. curvatus also has inclined limbs (Fürsich, 1974a; Chamberlain, 1977; Fillion & Pickerill, 1990). Arenicolites statheri has narrow, parallel and vertical limbs, and both A. statheri and A. naraensis have a thick wall lining (Fürsich, 1974a; Chamberlain, 1977; Fillion & Pickerill, 1990). Arenicolites sparsus typically lacks a wall lining but usually occurs only as paired openings on the tops of beds (Fürsich, 1974a). Arenicolites carbonaria consists of a small-diameter U-shaped tube with a very thin wall lining and funnel-shaped apertures (Fürsich, 1974a; Fillion & Pickerill, 1990). Arenicolites longistriatus is a U-shaped burrow that is subhorizontal after compaction and has longitudinal striations along the length of the burrow, most commonly at the base of the U-shaped tube (Rindsberg & Kopaska-Merkel, 2005).

  • The ichnotaxonomic status of some Arenicolites ichnospecies is currently debated. Recently, Mcllroy, Crimes, and Pauley (2005) and Callow, Mcllroy, and Brasier (2011) reexamined the type material of Arenicolites sparsus, the first ichnospecies established, and found that the depressions Salter (1857) interpreted as paired burrow apertures were in fact not paired and not connected together by a U-shaped tube. Arenicolites sparsus was reinterpreted as body fossils of small microbial mats and transferred to Beltanelliformis Menner in Keller & others, 1974 (McIlroy, Crimes, & Pauley, 2005; Callow, McIlroy, & Brasier, 2011). Menon and others (2015) later reinterpreted Beltanelliformis as a pseudofossil formed from fluid injection through the sediment and, therefore, A. sparsus is also likely a pseudofossil.

  • Arenicolites is considered a dwelling or suspension-feeding burrow of annelid worms or small arthropods (e.g., Hakes, 1976; Bromley & Asgaard, 1979; Fillion & Pickerill, 1990). Arenicolites has been reported mostly from shallow marine deposits, but freshwater-aquatic and deep-marine deposits have been reported as well (e.g., Crimes & others, 1977; Bromley & Asgaard, 1979; Fillion & Pickerill, 1984; Hasiotis, 2002, 2004, 2008; Ash & Hasiotis, 2013). Arenicolites ranges from the early Cambrian to recent (e.g., Crimes, 1987, 1992) with problematic specimens reported from the Neoproterozoic (e.g., Fillion & Pickerill, 1990).

  • ARENICOLITES CARBONARIA (Binney, 1852)
    Figure 6.3–6.4

  • Material.—IBGS PJ-M-003: multiple specimens, Spence Shale float, Cataract Canyon.

  • Diagnosis.—Vertical, U-shaped tubes expressed in plan view as paired depressions (concave epirelief) with small diameter limbs and funnel-shaped apertures in cross section (Fürsich, 1974a; Fillion & Pickerill, 1990).

  • Description.—Specimens are preserved as paired openings in concave epirelief. Apertures 1.0–2.2 mm wide, 0.7–1.1 mm deep, with variable spacing 0.7–4.1 mm. Burrow limbs are narrower than burrow openings and range 0.2–0.9 mm wide. Most limbs lack or have very thin wall linings (> 0.1 mm).

  • Occurrence.—Gray, massive, peloidal carbonate wackestone and packs tone to mudstone with thin continuous and discontinuous laminations of tan to brown, very fine-grained siliciclastic sandstone and siltstone. Soft-sediment deformation is present, but limestone and laminations are extensively bioturbated (ii3–4).

  • Associated ichnotaxa.Conichnus conicus.

  • Discussion.—Specimens were assigned to Arenicolites carbonaria due to their small, paired depression morphology. These specimens occur on the same slab as the Conichnus conicus described herein, but in a horizon ∼ cm below the Conichnus specimens (Fig. 6.3). Since Arenicolites is usually indicative of shallow-marine settings (e.g., Fillion & Pickerill, 1990) and due to its close vertical proximity to the Conichnus, sample IBGS PJ-M-003 is interpreted to have been deposited in shallower water and/or higher energy settings than the other Spence Shale ichnofossils. Though complete U-shaped tubes connecting surficial depressions are not visible in areas of the massive limestone in cut slabs, one complete U-shaped tube and several partial tubes are visible in weathered sections connecting funnel shapes in the thin laminations of very fine sandstone and siltstone (Fig. 6.4).

  • Ichnogenus AULICHNITES Fenton & Fenton, 1937b

  • Type ichnospecies.—Aulichnites parkerensis Fenton & Fenton, 1937b.

  • Diagnosis.—Convex epirelief, bilobate, ribbon trail with a medial furrow separating lobes; lower surface may show a unilobate, convex-downward shape, or as concave furrows with convex medial ridge (Fenton & Fenton, 1937b; Fillion & Pickerill, 1990).

  • Discussion.—Aulichnites is similar to other convex bilobate epirelief ichnotaxa, including Gyrochorte, Olivellites Fenton & Fenton, 1937c, Psammichnites Torell, 1870, and Scolicia de Quarterages, 1849. Aulichnites and Gyrochorte are composed of paired convex ridges with a medial furrow (epirelief) ; however, the ridges of well-preserved Gyrochorte have a biserial-plaited ornamentation (Häntzschel, 1975), and may occur as vertical stacks of bilobate, concave-down spreite (Heinberg, 1973, p. 231, fig. 6). Aulichnites may be similar to poorly preserved Gyrochorte specimens that lack the plaited ornamentation, like those illustrated by Heinberg (1973). D'Alessandro and Bromley (1987) and Mángano, Buatois, and Rindsberg (2002) synonymized Aulichnites under Psammichnites after interpreting Aulichnites to be a preservational variant of Olivellites, which was also considered a junior synonym of Psammichnites. Though similar, Aulichnites and Olivellites were established separately by Fenton and Fenton (1937b, 1937c) due to the presence of a medial furrow or ridge in epirelief, respectively. Chamberlain (1971) synonymized Aulichnites under Scolicia with no reason given; however, Häntzschel (1975) rejected the Chamberlain (1971) synonymy and most subsequent authors (e.g., Hakes, 1976, 1977; Fillion & Pickerill, 1990) have followed Häntzschel's rejection.

  • Aulichnites is interpreted as the locomotion or grazing trail of a gastropod (Fenton & Fenton, 1937b; Fillion & Pickerill, 1990); however, some authors considered them to have been produced by xiphosurids (i.e., horseshoe crabs; Yochelson & Schindel, 1978; Chisholm, 1985; Fillion & Pickerill, 1990). Aulichnites occurs in shallow- and deep-marine as well as brackish water deposits (e.g., Fenton & Fenton, 1937b; Fillion & Pickerill, 1990). Aulichnites ranges from the Ediacaran to recent (e.g., Häntzschel, 1975; Narbonne & Aitken, 1990; Crimes, 1992; Buatois & Mángano, 1993b; Jenkins, 1995; MacNaughton, 2003).

  • AULICHNITES isp.
    Figure 6.5

  • Material.—IBGS PJ-M-019: one specimen, Miner's Hollow.

  • Diagnosis.—Paired concave furrows separated by a medial ridge in hyporelief.

  • Description.—Bilobate, concave furrows separated by convex ridge in hyporelief, 7.0 mm long and 3.4 mm wide.

  • Occurrence.—Gray, laminated, siliciclastic or calcareous silty shale. Laminations are continuous with very little bioturbation occurring to disrupt them, indicating an ii2. The exposed bedding plane has extensive bioturbation with overprinting, indicating a (BPBI)3–4.

  • Associated ichnotaxa.—Dimorphichnus isp., Lockeia siliquaria, Phycosiphon incertum, Protovirgularia cf. pennatus, and Treptichnus vagans.

  • Discussion.—This specimen occurs as a short bilobate trail of paired concave ridges. The Aulichnites specimen terminates at a unilobate convex hyporelief mound. This mound is likely a short amygdaloidal (almond-shaped) resting trace assigned herein as Lockeia siliquaria. We consider the Aulichnites and Lockeia specimens to represent a compound trace (sensu Bertling & others, 2006) with the tracemaker having burrowed through the sediment (Aulichnites) and then stopped to rest (L. siliquaria).

  • Ichnogenus BERGAUERIA Prantl, 1945

    Figure 7.

    Plug-shaped ichnofossil specimens from the Spence Shale. 1–2, Bergaueria hemispherica in convex hyporelief, (1) profile view and (2) upper plan view with coarse infill (circle) and a trilobite pygidium fragment (arrow), KUMIP 314231; 3–4, Bergaueria hemispherica in concave epirelief (3) and cast (4), IBGS PJ-M-021, Miner's Hollow; 5–6, Bergaueria hemispherica in concave epirelief (5) and cast (6) with asymmetric shape and transverse constrictions, IBGS PJ-M-029, Miner's Hollow; scales in cm.

    f07_01.jpg

    BERGAUERIA HEMISPHERICA Crimes & others, 1977
    Figure 6.3–6.6, Figure 7.1–7.6

  • Material.—KUMIP 314229, KUMIP 314231, IBGS PJ-M-020, IBGS PJ-M-021, IBGS PJ-M-029: one specimen each, Miner's Hollow; IBGS PJ-M-025, two specimens, float from Cataract Canyon.

  • Diagnosis.—Vertical, hemispherical, plug-shaped burrow lacking shallow, central depression at apex of the burrow (Crimes & others, 1977).

  • Description.—Circular to elliptical plug-shaped depressions (concave epirelief) and mounds (convex hyporelief), diameter 15.8–40.8 mm, 4.3–17.2 mm thick, and diameter/thickness (D/T) ratio 1.5–3.5. Some epirelief specimens have transverse, ledge-like constrictions along burrow wall, hyporelief specimens have smooth walls; and lack both radial ridges and a central depression (hyporelief) or knob (epirelief) on base.

  • Occurrence.—Gray (weathered to brown), laminated calcareous or siliciclastic silty shale and sandy shale.

  • Associated ichnotaxa.—Cruziana barbata, Planolites annularis, Rusophycus carbonarius, Sagittichnus lincki, and Teichichnus cf. nodosus.

  • Discussion.—The majority of specimens assigned to this ichnospecies occur in concave epirelief on individual slab samples. Bergaueria hemispherica specimens have smooth rounded bases that lack small knobs (concave epirelief) or depressions (convex hyporelief) characteristic to other Bergaueria ichnospecies (Pemberton, Frey, & Bromley, 1988). Ledgelike constrictions (Fig. 7.3–7.6) occur transversely along burrow wall and are similar to constrictions associated with Conostichus. Crimes and others (1977) noted a similar concentric ornamentation and suggested it represents mudrich laminations not related to tracemaker morphology. Specimens lack radial ridges that would justify assignment to B. radiata, B. perata, or even Conostichus. One B. hemispherica specimen (Fig. 7.5–7.6) does bear a strong resemblance to Conostichus broadheadi due to the presence of a well-developed conical shape and narrow apical disc but lacks the distinctive longitudinal fluting. The 1.5–3.5 D/T ratios fit with those suggested by Pemberton, Frey, and Bromley (1988) for Bergaueria.

  • BERGAUERIA aff. PERATA (Prantl, 1945)
    Figure 8.1

  • Material.—IBGS PJ-M-026: one specimen (part and counterpart), Miner's Hollow.

  • Diagnosis.—Smooth walled, unlined or thinly lined, cylindrical mounds in convex hyporelief; faint ridges present radiating from a central depression may be present; diameter is generally equal to or greater than thickness (height) (Prantl, 1945; Pemberton, Frey, & Bromley, 1988).

  • Description.—Smooth, low relief depression (mound in convex hyporelief), 10.0 mm in diameter, 1.2 mm thick (height), and has a diameter-thickness ratio (D/T) of 8.33. No discernable radial ridges or central depressions are present.

  • Occurrence.—Gray (weathered to brown), calcareous silty shale.

  • Associated ichnotaxa.—None.

  • Discussion.—Only one specimen was collected and described from the Spence Shale. Bergaueria perata Prantl, 1945, was erected for unlined or thinly lined plug-shaped ichnofossils that may have diameters significantly greater than its thickness. Shallow and smooth B. aff. perata specimens may also be similar to Bergaueria sucta Seilacher, 1990–smooth, low relief, disclike basal impressions of actinians in laterally repeated sets that indicate lateral movement or creeping (Jensen, 1997). The smooth low relief and high D/T ratio is suggestive of an affinity to B. sucta, but the lack lateral repetition would preclude assignment as such.

  • Ichnogenus CONICHNUS Männil, 1966

  • Type ichnospecies.—Conichnus conicus Männil, 1966.

  • Diagnosis.—Short to long, vertical, cone-shaped to subcylindrical burrows with smooth, rounded base or randomly oriented papillalike protuberances on base; burrow infill maybe unstructured or have V-shaped laminations (Männil, 1966; Pemberton, Frey, & Bromley, 1988).

  • Discussion.—Conichnus is similar to several plug-shaped ichnofossils. Pemberton, Frey, and Bromley (1988) conducted a detailed review of 15 plug-shaped ichnogenera and synonymized them together into five ichnogenera; Astropolichnus, Bergaueria, Conichnus, Conostichus, and Dolopichnus. Conichnus is a conical to subcylindrical burrow with smooth walls and rounded base (C. conicus), but the base may have protuberances (C. papillatus) (e.g., Männil, 1966; Frey & Howard, 1981; Pemberton, Frey, & Bromley, 1988). Conostichus is distinguished by transverse constrictions and longitudinal fluting of the burrow wall, a basal apical disc, and a burrow diameter approximately twice its height. Bergaueria is characterized by a cylindrical to hemispherical shape, thick to thin wall linings, a central depression and/or radial ridges on the base, and a diameter twice its height. Dolopichnus is distinguished by a larger size, a central cylindrical core typically with coarser infill, bulb-shaped terminations in some, and a diameter roughly one quarter its height. Astropolichnus is a short cylinder with a diameter over three times its height, radial ridges, and a central core (Pemberton, Frey, & Bromley, 1988).

  • Conichnus is commonly interpreted as dwelling or resting traces of actinians (e.g., sea anemones) (e.g., Pemberton, Frey, & Bromley, 1988; Mangano & others, 2002). Most Conichnus are reported from shallow-marine deposits and tidal deposits (e.g., Frey & Howard, 1981; Hiscott, James, & Pemberton, 1984; Mángano & others, 2002). Conichnus ranges from the early Cambrian to recent (e.g., Curran & Frey, 1977; Hiscott, James, & Pemberton, 1984; Jackson, Hasiotis, & Flaig, 2016).

  • CONICHNUS CONICUS Männil, 1966
    Figure 8.2–8.4

  • Material.—IBGS PJ-M-003: 12 specimens, Spence Shale float, Cataract Canyon.

  • Diagnosis.—Short cone- to plug-shaped depression with smooth, rounded bottom, some penetrated by vertical tube.

  • Description.—Short, plug-shaped depression with smooth base filled with massive, gray calcareous mudstone; 4–10 mm in diameter and 1–4 mm deep. Central-plug diameter 2.6–2.9 mm.

  • Occurrence.—Tan to brown, very fine-grained sandstone with ripple marks above a layer of gray peloidal carbonate wackestone and packstone to mudstone with thin, tan to brown silty to sandy laminations and soft-sediment deformation; however, no Conichnus specimens are present in the lower layer.

  • Associated ichnotaxa.—Arenicolites carbonarius.

  • Discussion.—Specimens were assigned to C. conicus for their small, pluglike morphology with smooth, rounded bottoms and lack of basal protuberances (Fig. 8.2–8.4). Conichnus conicus specimens occur on the same sample with Arenicolites carbonarius but are restricted to a higher layer. Most C. conicus specimens occur close to another specimen and falsely appear as openings to U-shaped burrows (e.g., Arenicolites or Diplocraterion, Fig. 8.2). A cut section of one specimen revealed a massive, carbonate mudstone infill penetrated by a central vertical tube (e.g., possible Skolithos; Fig. 8.3–8.4) suggesting that some C. conicus may be composite traces (i.e., two or more unrelated ichnotaxa occurring within each other; sensu Bertling & others, 2006).

  • Figure 8.

    Plug-shaped ichnofossil specimens from the Spence Shale (continued). 1, Bergaueria aff. perata in convex hyporelief, IBGS PJ-M-026, Miner's Hollow; 2, plan view of Conichnus conicus in concave epirelief, IBGS PJ-M-003, Cataract Canyon; 3—4, Cross-sections of C. conicus and Skolithos-like vertical tubes (arrows); scales in cm.

    f08_01.jpg

    Ichnogenus CRUZIANA d'Orbigny, 1842

  • Type ichnospecies.—Cruziana rugosa d'Orbigny, 1842, by subsequent designation in Miller (1889).

  • Diagnosis.—Elongate, bilobate, ribbonlike furrows with medial ridges (concave epirelief) or grooves (convex hyporelief): furrows commonly covered by herringbonelike, transverse, or longitudinal striations (Crimes, 1970a, 1970b; Seilacher, 1970; Häntzschel, 1975).

  • Discussion.—Seilacher (1970) united both bilobate long furrows and short excavations (=Rusophycus) under Cruziana due to similar striation patterns (interpreted as scratch marks) attributed to the same organism, trilobites; however, this proposal was rejected by numerous authors (e.g., Crimes, 1970a, 1970b, 1975; Fillion & Pickerill 1990; Pickerill, 1995; Jensen, 1997) due to significant morphologic differences between the two ichnogenera. Bromley and Asgaard (1979) included ribbonlike Isopodichnus Bornemann, 1889, under Cruziana because the two ichnogenera differ only in accessory features (e.g., size), which is suggested for use only in ichnospecific designation (sensu Fürsich 1974b). The Cruziana Isopodichnus synonymy, though rejected by Hakes (1985), Pollard (1985), and Seilacher (1985), is still followed by most authors. Crimes (1970b) noted that Cruziana can grade into other ichnogenera (e.g., Diplichnites, Diplopodichnus Brady, 1947, and Rusophycus) and that the V-shaped striations open in the direction of movement as with Diplichnites.

  • Cruziana is commonly interpreted as a surficial to shallow deposit-feeding, dwelling, grazing, locomotion, or predation trace (e.g., Crimes, 1970a, 1970b; Seilacher, 1970; Zonneveld & others, 2002; Gingras & others, 2007). Most Cruziana have been interpreted as the product of trilobites but other tracemakers have been suggested; nontrilobite arthropods (e.g., horseshoe crabs, branchiopods, aglaspidids), or even some vertebrates (e.g., Seilacher, 1970; Fisher, 1978; Shone, 1978, 1979; Bromley & Asgaard, 1979, Pollard, 1985). Cruziana has been reported in deep- and shallow-marine and continental deposits (e.g., fluvial, lacustrine, and brackish) (e.g., Crimes, 1970a, 1970b; Bromley & Asgaard, 1979; Seilacher, 1985; Pillion & Pickerill, 1990; Pickerill, 1995). Cruziana ranges from the early Cambrian to the Cretaceous (e.g., Crimes, 1987, 1992; Mangano & others, 2002; Hasiotis, 2012).

  • Figure 9.

    Cruziana specimens from Spence Shale. 1–2, Cruziana barbata in concave epirelief, KUMIP 314229, Miner's Hollow; 1, two specimens of C. barbata overlapped in opposite directions (arrows); 3, Rusophycid C. problematica with small Rusophycus carbonarius in convex hyporelief, IBGS PJ-M-007, Miner's Hollow; 4, Cruziana problematica with Lockeia siliquaria (arrow) and Monomorphichnus cf. multilineatus (circle) in convex hyporelief, KUMIP 314228, Miner's Hollow; scale in cm.

    f09_01.jpg

    CRUZIANA BARBATA Seilacher, 1970
    Figure 9.1–9.2

  • Material.—KUMIP 314229; eight specimens, Spence Shale, Miner's Hollow, Wellsville Mountains, Utah, USA.

  • Diagnosis.—Small to medium, straight to curved, bilobate ribbonlike furrow with medial ridge and curved V-shaped striations angled ∼60° (Seilacher, 1970; Legg, 1985).

  • Description.—Bilobate, concave epirelief, ribbon trails; 27.0– 94.4 mm long and 9.8–11.9 mm wide. Curved striations are visible in several specimens and have a V-shaped angle 142–163°.

  • Occurrence.—Greenish gray (weathered to brown) calcareous, micaceous silty shale.

  • Associated ichnotaxa.—Bergaueria hemispherica, Planolites annularis, and Rusophycus carbonarius.

  • Discussion.—Specimens assigned to C. barbata partly crossover other C. barbata specimens on the same sample along their lengths causing some lobes to be lost and give the appearance of a trilobate form, but the specimens can be differentiated, as the V-shaped striations are oriented opposite to the overlapping furrow (Fig. 9.1). Several trilobite pygidia are present on the sample; however, their widths are greater than the widths of the C. barbata specimens indicating that those trilobites were not the producers (sensu Fortey & Seilacher, 1997). Spence Shale specimens of C. barbata are significantly smaller (∼1 cm) than most previously recorded specimens (∼3–9 cm; Legg, 1985; Orłowski, 1992). The decreased size is likely the result of lower available oxygen near the sediment-water interface (e.g., G arson & others, 2012).

  • CRUZIANA PROBLEMATICA (Schindewolf, 1921)
    Figure 9.3–9.4, Figure 10.1–10.5, 16.5

  • Material.—KUMIP 204523 A (part) and B (counterpart): 31 specimens, Miners Hollow; KUMIP 314228; two specimens, Miner's Hollow; IBGS PJ-M-007: five specimens, Miners Hollow; IBGS PJ-M-016: two specimens, Miners Hollow; IBGS PJ-M-017: one specimen, Miners Hollow float.

  • Diagnosis.—Small to large, straight to curved, bilobate ribbonlike furrow with medial ridge and transverse striations (Bromley & Asgaard, 1979; Fillion & Pickerill, 1990; Jensen, 1997).

  • Description.—Concave or convex, bilobate, ribbonlike burrows with a medial ridge (epirelief) or furrow (hyporelief) and transverse striations; Burrows 11.4–95.1 mm long and 8.0–15.4 mm wide. Striation V-shaped angle ∼80° but some range from 145–160°. Burrow paths are typically slightly curved to straight, but several burrows are highly curved and overlap or crosscut each other.

  • Occurrence.—Greenish gray to gray laminated, calcareous silty to sandy shale; sometimes weathered to brown or brownish yellow.

  • Associated ichnotaxa.—Lockeia siliquaria, Monomorphichnus lineatus, M. cf. multilineatus, Planolites beverleyensis, P. montanus, Rusophycus carbonarius, Rusophycus cf. cerecedensis, Treptichnus bifurcus, and T. pedum.

  • Discussion.—Bromley and Asgaard (1979) placed ribbonlike Isopodichnus Bornemann, 1889, under Cruziana problematica because size was not enough to warrant a separate ichnogenus. Some authors, however, retain Isopodichnus for use as a salinity indicator in fresh- and brackish-water settings (e.g., Hakes 1985; Pollard 1985; Seilacher 1985, 2007). Bromley and Asgaard (1979) also noted that Isopodichnus was reported from marine deposits by Alpert (1976a) and Trewin (1976), thus, making retention of Isopodichnus as a salinity indicator invalid. Jensen (1997) attempted to distance the ichnospecies from the common interpretation as a salinity indicator by placing it under a resurrected name, Cruziana tenella (Linnarsson, 1871) (for discussion, see Jensen, 1997). Reassignment of C. problematica to C. tenella has been accepted by some authors (e.g., MacNaughton & Narbonne 1999; Jensen, Droser, & Heim, 2002; Zonneveld & others, 2002; Sadlok, 2010), but rejected by others for nomenclatural stability (e.g., Mángano & others, 2002; Schatz & others, 2011). We reject the renaming of C. problematica to C. tenella in favor of nomenclatural stability, even though the use of ichnotaxa as environmental stress indicators is not valid to establish, rename, or retain ichnotaxa.

  • Cruziana problematica specimens show some meandering, suggesting they were produced via grazing, and are noticeably larger than R. carbonarius (Fig. 9.3–9.4, Fig. 10.1–10.5). The average width of C. problematica is 10 mm, whereas R. carbonarius averages ∼5 mm wide. The width difference suggests that C. problematica tracemakers were not the same as the tracemakers of R. carbonarius (sensu Fortey & Seilacher, 1997), which could be juveniles of the adult form (Cruziana producers). Cruziana problematica and C. problematica-sized Rusophycus specimens do not occur together on KUMIP 204523, although one association does occur on KUMIP 314228 (see Fig. 9.4). The specimens of Cruziana problematica on KUMIP 204523 co-occur with Rusophycus carbonarius, Planolites montanus, and Treptichnus bifurcus and were likely not constructed at the same time and may have been affected by sudden changes in available oxygen or nutrients. Rusophycus carbonarius specimens crosscut both C. problematica and other R. carbonarius (Fig. 10.1), whereas Cruziana problematica specimens only crosscut each other (Fig. 10.3). Planolites montanus crosscuts both C. problematica and R. carbonarius (Fig. 10.4). The crosscutting relationships suggest that C. problematica were constructed and abandoned first, followed by R. carbonarius, and then finally, P. montanus. The T. bifurcus specimen was constructed sometime after the C. problematica as the latter was cross cut by the former (Fig. 10.5), but its placement in the aforementioned crosscutting timeline is unknown because the T. bifurcus specimen has no interaction with any other specimen.

  • Ichnogenus DIMORPHICHNUS Seilacher, 1955a

  • Type ichnospecies.—Dimorphichnus obliquus Seilacher, 1955a.

  • Diagnosis.—Asymmetrical trackways with two types of impressions, typically of equal width; (1) long, thin, straight to sigmoidal striations; and (2) short, punctate to elliptical impressions at end of long striations; both types occur oblique to direction of movement (Seilacher, 1955a; Fillion & Pickerill, 1990).

  • Discussion.—Seilacher (1955a) named Dimorphichnus for oblique sets of elongate striations with punctate impressions produced by trilobites. The movement of the Dimorphichnus tracemaker was oblique to fully sideways with the short, punctate impressions formed by one set of legs acting as a holdfast to keep the tracemaker in place, while the sigmoidal striations were formed by the other set of legs sweeping through the medium returning to their starting position (Seilacher, 1955a, 2007). Crimes (1970a, 1970b) suggested the oblique to sideways orientation of Dimorphichnus was due to increased current energy forcing the tracemaker to reorient itself to remain stable while moving or grazing. After Monomorphichnus was described by Crimes (1970b), Seilacher (1985) argued that Monomorphichnus was a junior synonym of Dimorphichnus and the Monomorphichnus holotype contained punctate impressions consistent with Dimorphichnus. Most authors have rejected this suggestion and maintain both as separate ichnogenera (e.g., Walter, Elphinstone, & Heys, 1989; Fillion & Pickerill, 1990; Orłowski, 1992; Jensen, 1997; Hofmann & others, 2012). More recently, Seilacher (2007) proposed that Dimorphichnus and Monomorphichnus should be considered as a behavioral and preservational variant of Diplichnites, respectively. Jensen (1997) and Hofmann and others (2012) suggested that Dimorphichnus and Monomorphichnus should remain separate due to each representing a separate behavior.

  • Dimorphichnus is interpreted as a locomotive, deposit feeding, or grazing trace (Seilacher, 1955a, 1985; Crimes, 1970b; Fillion & Pickerill, 1990). Proposed producers of Dimorphichnus include marine and continental arthropods (e.g., trilobites, decapods, centipedes, millipedes) (Fillion & Pickerill, 1990). Dimorphichnus has been reported from shallow-marine, deep-marine, and continental deposits (e.g., alluvial, lacustrine, and eolian) (e.g., Seilacher, 1955a; Fillion & Pickerill, 1990). Dimorphichnus ranges from the Cambrian to recent (e.g., Crimes, 1970b).

  • Figure 10.

    Cruziana problematica and Dimorphichnus specimens from the Spence Shale. 1–5, Cruziana problematica, KUMIP 204523A and B, Miner's Hollow; 1–2, Cruziana problematica with several Rusophycus carbonarius (arrows), convex hyporelief; 3, Cruziana problematica with transverse striations, crosscut by Planolites montanus and R. carbonarius; 4, overlapping C. problematica, with grazinglike scribble paths; 5, convex hyporelief of C. problematica with several R. carbonarius (black arrows) and Treptichnus bifurcus (white arrows) in both convex and concave hyporelief; 6, Dimorphichnus isp., rakes (black arrows) and pusher (white arrow), in convex hyporelief, IBGS PJ-M-019, Miner's Hollow; scale in cm.

    f10_01.jpg

    DIMORPHICHNUS isp.
    Figure 10.6

  • Material.—IBGS PJ-M-019; one specimen, Miner's Hollow; IBGS PJ-M-024: one specimen, Miner's Hollow.

  • Diagnosis.—Small, thin, laterally repeated sets of elongate, sigmoidal striations (convex hyporelief) with separate blunt, ovoid to circular mounds occurring near end of elongate striations.

  • Description.—Specimens consist of thin convex ridges and separate ovoid to punctate mounds near the ridge ends. Trackways 19.6 mm long, 7.1 mm wide. Sigmoidal striations 4.0–6.2 mm long, 0.2–0.4 mm wide. The blunt mounds 0.8–2.0 mm long, 0.4–0.6 mm wide.

  • Occurrence.—Gray, laminated, siliciclastic or calcareous silty shale. Laminations are continuous with very little bioturbation occurring to disrupt them, indicating an ii2. The exposed bedding plane has extensive bioturbation with some overprinting, indicating a BPBI 3–4.

  • Associated ichnotaxa.—Aulichnites isp., Lockeia siliquaria, Phycosiphon incertum, Planolites annularis, P. montanus, Protovirgularia cf. pennatus, Rusophycus carbonarius, Sagittichnus lincki, and Treptichnus vagans.

  • Discussion.—The specimens of Dimorphichnus are very diminutive in size with sigmoidal striation. Pusher mound widths are <1.0 mm (Fig. 10.6). The Dimorphichnus isp. on IBGS PJ-M-019 does not crosscut any recognizable traces. The extensive bioturbation of the base of IBGS PJ-M-019 makes identification of ichnotaxa difficult and suggests a more oxygenated environment than the shallower laminations yielding Phycosiphon incertum.

  • Ichnogenus DIPLICHNITES Dawson, 1873

  • Type ichnospecies.—Diplichnites aenigma Dawson, 1873.

  • Diagnosis.—Simple trackways of punctate to elongate track impressions in parallel track rows; track impressions closely and regularly spaced, and normal or oblique to trackway axis (Häntzschel, 1975; Briggs, Rolfe, & Brannan, 1979; Fillion & Pickerill, 1990).

  • Discussion.—Originally interpreted as trails of large myriapods or annelids by Dawson (1873), recent authors have used Diplichnites to describe smaller-scaled trackways thought to be produced by trilobites (Fillion & Pickerill, 1990). Briggs, Rolfe, and Brannan (1979) suggested that Diplichnites be restricted to continental arthropod trackways because they noted that workers were departing from the original diagnosis of Diplichnites as a continental trackway and suggested using some junior synonyms of Diplichnites from Osgood (1970) to place the trilobite-produced trackways.

  • Nine ichnospecies are currently recognized within the literature (e.g., Buatois & others, 1998; Smith & others, 2003): Diplichnites aenigma Dawson, 1873; D. binatus Webby, 1983; D. cuithensis Briggs, Rolfe, & Brannen, 1979; D. govenderi Savage, 1971; D. gouldi (Gevers in Gevers & others, 1971); D. incertipes (Matthew, 1910); D. minimus Walter & Gaitzsch, 1988; D. minor (Matthew, 1910), and D. triassicus (Linck, 1943). Track orientation and shape, number of tracks per track series, and number of track series per track row are generally used to differentiate ichnospecies (e.g., Savage, 1971; Trewin & McNamara, 1995). Trewin and McNamara (1995) divided D. gouldi into three morphotype end-members (types A, B, and C) based on trackway width and the number of tracks per track series.

  • Diplichnites is generally interpreted as a locomotion trace of trilobites but other arthropods, including myriapods, and some annelids have been suggested (e.g., Dawson, 1873; Osgood, 1970; Briggs, Rolfe, & Brannen, 1979). Diplichnites is found in shallowand deep-marine, and continental deposits (e.g., Crimes & others, 1977; Fillion & Pickerill, 1990; Crimes & Fedonkin, 1994). Deep-marine Diplichnites is mostly reported from the lower and middle Cambrian and only rarely after the Cambrian (Pickerill, 1981; Crimes & Fedonkin, 1994). Diplichnites ranges from the Cambrian to recent (e.g., Briggs, Rolfe, & Brannen, 1979; Crimes, 1987, 1992; Fillion & Pickerill, 1990; Hasiotis, 2012).

  • DIPLICHNITES GOULDI (Gevers in Gevers & others, 1971)
    TYPE A Trewin & McNamara, 1995
    Figure 11.1–11.3

  • Material.—KUMIP 204522: one specimen, Antimony Canyon; IBGS PJ-M-011: one specimen, Spence Tongue of the Lead Bell Shale, Oneida Narrows, Bear River Range, Idaho, USA; IBGS PJ-M-014 (part and counterpart) and IBGS PJ-M-015: one specimen, Miner's Hollow.

  • Diagnosis.—Paired rows of punctate to ellipsoidal or elongated straight impressions oriented perpendicular or oblique to trackway axis; track series consist of 5–9 tracks in opposition. Within track rows, multiple sets of track impressions may overlap previous sets (Trewin & McNamara, 1995; Buatois & others, 1998; Smith & other, 2003).

  • Description.—Trackways 32.1–49.0 mm long; outer trackway 10–15 mm wide, inner trackway 8.4–9.5 mm wide. Punctate to ellipsoidal tracks 2–4 mm wide, spaced 2.5–4.0 mm apart. Specimens with overlapping track series, overlap occurs by 2–3 tracks, overlap distance 1.5–2.9 mm.

  • Occurrence.—Two lithologies: (1) gray to dark gray (weathered to tan), very fine to fine carbonate sand to silty shale; and (2) pale greenish gray, mica-rich, silty to sandy shale. Thin to thick laminations are present, but are unbroken or have rare traces in slab samples (ii1–2). Bedding plane is only disrupted by D. gouldi (BPBI 2).

  • Associated ichnotaxa.—Planolites montanus and Treptichnus vagans.

  • Discussion.—Diplichnites gouldi was originally described by Gevers in Gevers and others (1971) for paired, parallel rows of punctate to ellipsoidal track impressions under the name Arthropodichnus gouldi. Gevers (1973) changed the name from Arthropodichnus to Beaconichnus since Arthropodichnus was already proposed for another ichnogenus. Bradshaw (1981) transferred Beaconichnus gouldi into Diplichnites as D. gouldi. Trewin & McNamara (1995) recognized three end-members (types A, B, and C) with material assigned to D. gouldi based on trackway widths and tracks per series. Buatois and others (1998), however, considered that D. gouldi type A did not belong in Diplichnites and viewed it as a form of Umfolozia Savage, 1971, while retaining D. gouldi types B and C. Smith and others (2003) suggest retaining all three end-members of D. gouldi, with which we agree.

  • Häntzschel (1975) placed Acripes Matthew, 1910, within Diplichnites due to similar morphology. Miller (1996) reviewed type material of Acripes and confirmed its placement in Diplichnites but made no reference or recommendation on whether all three Acripes ichnospecies should remain valid under Diplichnites. Some authors have included A. incertipes, A. leavitti, and A. minor as valid ichnospecies within Diplichnites (e.g., Keighley & Pickerill, 1998; Smith & others, 2003). Keighley and Pickerill (1998) recommended that A. incertipes (Matthew, 1910, pl. III, fig. 1–2) should not be included in Diplichnites due to significantly different track impression shapes of each track row similar to Dimorphichnus, Petalichnus, and Ptilichnus.

  • The ichnospecies of Matthew (1910), Acripes incertipes (sensu stricto; plate III, fig. 1), A. leavitti, and A. minor, are morphologically almost identical to Diplichnites gouldi. Each ichnospecies are paired, parallel trackways with punctate, opposite track impressions in series that may overlap and are differentiated primarily by size. We suggest that they should be grouped under a single ichnospecies, Diplichnites gouldi, as it: (1) has the most similar morphology to Acripes; (2) is the most commonly used in the literature; and (3) would help stabilize the nomenclature regarding Diplichnites.

  • Specimens assigned to Diplichnites gouldi type A consist of small, punctate to ellipsoidal track impressions. Most D. gouldi type A specimens occur with specimens of Planolites montanus, but one specimen is present alongside Monomorphichnus bilinearis as well as P. montanus (IBGS PJ-M-011). Most trackways are straight to gently curved, with the track series being most apparent in the curved sections. Some specimens have punctate tracks (Fig. 11.1); however, tracks are typically ellipsoidal and oriented ∼45–90° from the trace axis (Fig. 11.2–11.3). Several of the ellipsoidal-track specimens show track impressions of both track rows that are oriented parallel in a single direction suggesting bottom currents influenced the movement of the tracemakers (Trewin & McNamara, 1995; Smith & others, 2003; Seilacher, 2007).

  • Figure 11.

    Trackway ichnofossil specimens from the Spence Shale. 1, Diplichnites gouldi in concave epirelief, IBGS PJ-M-015, Miner's Hollow; 2, Diplichnites gouldi in convex hyporelief, IBGS PJ-M-014, Miner's Hollow; 3, Diplichnites gouldi in concave epirelief, KUMIP 204522, Antimony Canyon; 4, Diplichnites cf. bitiatus with paired impressions (arrows) in convex hyporelief, KUMIP 204521 A and B, Miner's Hollow; 5, close up of D. cf. govenderi track impressions with Protovirgularia cf. pennatus in convex hyporelief, KUMIP 204521 A and B; 6, Diplichnites cf. govenderi (white arrows) crosscut by D. cf. binatus (black arrow) in concave epirelief, KUMIP 204521 A and B. 1–4, 6,s Scale in cm; 5, scale in mm.

    f11_01.jpg

    DIPLICHNITES cf. BINATUS Webby, 1983
    Figure 11.4, 11.6

  • Material.—KUMIP 204521: one specimen (part and counterpart), Miner's Hollow.

  • Diagnosis.—Paired rows of thin, straight, elongated striations grouped in pairs or triplets oriented obliquely to the trackway axis; track impression morphology may be asymmetric (Webby, 1983; Buatois & others, 1998).

  • Description.—The left track row (relative to inferred tracemaker movement) is poorly preserved compared to the right track row. Trackway 153 mm long; outer trackway 23.8 mm wide, and inner trackway 18.4 mm wide. Thin, elongate striations 7.6–12.2 mm long, 0.6–1.6 mm wide, and spaced 1.2–5.9 mm apart. Track impressions oriented 45° from the central axis with a ∼90° V-shaped angle.

  • Occurrence.—Dark gray (weathered to tan) to pale greenish gray, very fine- to fine-grained carbonate silty shale.

  • Associated ichnotaxa.—Diplichnites cf. govenderi and Protovirgularia cf. pennatus.

  • Discussion.—The specimen assigned to Diplichnites cf. binatus occurs with other surficial arthropod trackways. The specimen is crosscut by a paired-row trackway with highly variable track impression morphologies, which ranges between punctate to apostrophelike to bifid to trifid morphologies of Keighley and Pickerill (1998) that is herein assigned to D. cf. govenderi. The D. cf. binatus specimen is poorly preserved and only one track row is clearly visible (Fig. 11.4), but shows a clear V-shape angle to indicate the tracemaker moved from right to left (relative to the image). Some of the elongate tracks occur in close pairs, which justify placement under D. binatus; however, some impressions are singular and others are in groups of three.

  • Diplichnites cf. binatus bears a resemblance to Pterichnus Hitchcock, 1865, as both ichnospecies have track impressions that are elongate and thin; however, D. cf. binatus commonly has asymmetrical impressions (Buatois & other, 1998), whereas the impressions of P. tardigradus are usually always symmetrical (Hitchcock, 1858, 1865; Gaillard & others, 2005). Minter, Mángano, and Caron (2012) suggested that Pterichnus and other similar V-forming trackways described by Hitchcock (1858, 1865) were actually undertracks and should be considered junior synonyms of Lithographus Hitchcock, 1858.

  • DIPLICHNITES cf. GOVENDERI Savage, 1971
    Figure 11.4–11.6

  • Material.—KUMIP 204521 A and B: two specimens, Miner's Hollow.

  • Diagnosis.—Paired rows of lunate to tapered to bifid track impressions oriented perpendicular to oblique to trackway axis; tracks may be opposite or staggered.

  • Description.—Trackways 47–190 mm long; outer trackways 30.6–45.3 mm wide, and inner trackway 9.5–15.8 mm wide. Lunate to tapered, bifid track impressions 3.6–12.2 mm long, 0.8–1.6 mm wide, and spaced 2.6–13.0 mm apart. Specimens lack overlapping series and form single-series track rows.

  • Occurrence.—Dark gray (weathered to tan) to pale greenish gray, very fine- to fine-grained calcareous silty shale. No visible bedding or laminations are present. Low to moderate bedding plane disruption by traces indicating BPBI 2.

  • Associated ichnotaxa.—Diplichnites cf. binatus and Protovirgularia cf. pennatus.

  • Discussion.—The specimen assigned to this ichnospecies has highly variable track impressions that make classification difficult; however, the closest ichnotaxa to which the Spence Shale material can be assigned are Diplichnites govenderi, Incisifex Dahmer, 1937, Lithographus or Permichnium Guthorl, 1934. The specimens differ from Incisifex because the track impressions are typically straight and elongate, whereas D. cf. govenderi have a mix of bifid, lunate, and elongate impressions (Häntzschel, 1975). The specimen differs from Lithographus because none of the track impressions have the trifid to J-shaped track impressions, whereas the holotype of D. govenderi (Savage, 1971, fig. 7A) shows lunate-shaped track impressions similar to those seen in the Spence Shale specimens. Permichnium differs from D. cf. govenderi as the track impressions are typically bifid and open either to the outside or inside of the trackway (Kramer & others, 1995), whereas D. cf. govenderi has multiple impression shapes.

  • A “quadrifid” track impression is present and is likely two tracks overprinting each other, composed of two bifid grooves that intersect near the outer margin of the trackway (Fig. 11.5). The quadrifid impression was likely produced via a two-part limb motion. First, an insertion of a bifid limb into the medium, which moved obliquely inward and to the posterior of the trackway, as indicated by a raised sediment mound near the end of the impression. Later, a second insertion that shifted obliquely inward toward the anterior of the trackway, which resulted in overlapping bifid impressions.

  • Figure 12.

    Gordia marnia specimens from the Spence Shale. 1, Gordia marnia in Banffia sp. BST carbon film in convex and concave hyporelief, IBGS LG-M-006; 2, Gordia marnia in convex hyporelief and endorelief, IBGS PJ-M-004, Miner's Hollow; 3, line drawing of G. marnia on IBGS LG-M-006; 4, line drawing of overlapping G. marnia burrows on IBGS PJ-M-004.

    f12_01.jpg

    Ichnogenus GORDIA Emmons, 1844

  • Type ichnospecies.—Gordia marnia Emmons, 1844, by original monotypy.

  • Diagnosis.—Smooth, winding but not meandering, unbranched, cylindrical burrows with common overcrossings and massive infill (Fillion & Pickerill, 1990; Wang & others, 2009).

  • Discussion.—Gordia was originally described and named for its resemblance to the freshwater hairworm, Gordius Linnaeus, 1758, but a poor definition caused some authors to view Gordia as nomen nudum (Emmons, 1844; Fillion & Pickerill, 1990). Hall (1847) provided a new description, which provided the diagnosis for Gordia as an ichnofossil (Fillion & Pickerill, 1990). Buatois and others (1998) suggested the synonymy of Haplotichnus Miller, 1889, under Gordia due to similar path irregularity and burrow overlap, even though Haplotichnus has frequent sharp bends in the burrow and rarely crosses itself. They considered the sharp bends in Haplotichnus to represent only a minor behavioral difference that yielded only an accessory feature (sensu Fürsich, 1974b; Buatois & others, 1998) and did not warrant separation. The sharp, irregular bends, however, are major architectural differences (sensu Hasiotis & Mitchell, 1993; Hasiotis, Mitchell, & Dubiel, 1993), as the sharp-angle bends and rare self-crossings are ichnotaxonomically significant at the ichnogeneric level. We, therefore, reject the synonymy of Haplotichnus within Gordia, and retain Haplotichnus as a separate ichnotaxon.

  • Gordia is commonly interpreted as a locomotion, depositfeeding, or grazing trace of annelid worms or other wormlike organisms, arthropods, or nematodes (e.g., Emmons, 1844, Buatois & Mángano, 1993b). Gordia is a one of most common faciescrossing ichnofossils known and has been reported from almost every depositional environment in deep and shallow marine, as well as, from estuarine, fluvial, and lacustrine deposits (e.g., Fillion & Pickerill, 1990; Buatois & Mángano, 1993b; Uchman, Kazakauskas, & Gaigalas, 2009; Jackson, Hasiotis, & Flaig, 2016). Gordia ranges from the Ediacaran to recent (Crimes & Anderson, 1985; McCann & Pickerill, 1988; Fillion & Pickerill, 1990; Wang & others, 2009; Hasiotis & others, 2012).

  • GORDIA MARNIA Emmons, 1844
    Figure 12.1–12.4

  • Material.—IBGS LG-M-006; one specimen; IBGS PJ-M-004; one specimen, Miners Hollow.

  • Diagnosis.—Thin, arcuate to winding burrows or trails with self-overcrossing patterns (De Gibert & others, 2000).

  • Description.—Winding burrows in convex epi- or hyporelief or in concave epirelief with multiple, arcuate, self-overcrossing trails. Burrows 4.2–38.4 mm long, 0.3–1.1 mm wide.

  • Occurrence.—Laminated, greenish gray (weathers to brown or yellowish brown), mica-rich siliciclastic shale with thin laminations of dark gray sandy shale. One specimen occurs in a BST carbonaceous film of Banffia sp. (J.B. Caron, personal communication, 2016).

  • Associated ichnotaxa.—None.

  • Discussion.—The occurrence G. marnia within the BST film (Fig. 12.1, 12.3) suggests that anoxic conditions were present for a period long enough to allow the Banffia sp. to decay into a BST carbon film before oxic or dysoxic conditions returned, allowing the Gordia tracemaker to feed off the remaining organic matter (sensu Wang & others, 2009; Garson & others, 2012). The numerous overlapping burrow segments on IBGS PJ-M-004 suggests a high concentration of detrital organics in the sediment (Fig. 12.2, 12.4).

  • Ichnogenus GYROPHYLLITES Glocker, 1841

  • Type ichnospecies.—Gyrophyllites kwassizensis Glocker, 1841.

  • Diagnosis.—Vertical to oblique shaft with numerous radiating club- to leaf-shaped tunnels or lobes on staggered levels and usually unbranched; each lobe may have been backfilled (Uchman, 1998).

  • Discussion.—Gyrophyllites is very similar to numerous rosetteshaped ichnofossils. Głuszek (1998) noted that Gyrophyllites bears a strong resemblance to Asterosoma Otto, 1854, when viewed in plan view where only one level of Gyrophyllites is viewed. Some authors have noted that Gyrophyllites looks similar to both Asterichnus Bandel, 1967, and Stelloglyphus Vialov, 1964 (e.g., Uchman, 1998, Le Roux, Nielson, & Henríquez, 2008). Similarities between Atollites Maas, 1902, and Gyrophyllites have been noted as both have radiating lobes and a theorized helical structure (Seilacher, 1977; Serpagli, 2005). However, the lobe terminations in Atollites are more spherical than club shaped or straight compared to Gyrophyllites. Lorenzinia Gabelli, 1900, is composed of radiating burrows with a large flat central area separating the inner burrow terminations and has no apparent central shaft (Häntzschel, 1975). Fürsich and Bromley (1985) reinterpreted Dactyloidites Hall, 1886, and remarked on its superficial similarity to Gyrophyllites and other rosette ichnofossils, but noted that Dactyloidites contained radial spreiten. The figures in Fürsich and Bromley (1985, fig. 7, 8, & 10), however, show vertically to sub vertically stacked spreiten with the exception of D. asterioides. Though commonly illustrated as a three-dimensional helical structure (e.g., Häntzschel, 1975, p. 66, fig. 40.2b), Gyrophyllites is thought by some to be a rosette trace occurring in multiple stories, with each restricted to a single bedding plane and connected by a central tube (e.g., Fürsich & Kennedy, 1975; Le Roux, Nielson, & Henriquez, 2008; Strzebonski & Uchman, 2015).

  • Gyrophyllites is interpreted as the feeding burrow system of a wormlike deposit feeder, such as polychaete and echiuran worms (e.g., Chamberlain, 1975; Fürsich & Kennedy, 1975; Mángano, Buatois, & Muñiz Guinea, 2005; Le Roux, Nielson, & Henriquez, 2008; Strzebonski & Uchman, 2015). Fürsich and Kennedy (1975) suggested that Gyrophyllites was produced preferentially in silty and clayey layers as the tracemaker mined sediment for food and stopped excavation when sand-rich layers were encountered. Gyrophyllites is most commonly reported from deep-marine flysch and fan overbank deposits but has also been reported from shallow-marine deposits (e.g., Wetzel & Uchman, 1997; Uchman, 1998; Seilacher, 2007; Strzebonski & Uchman, 2015). Gyrophyllites ranges from the Cambrian to Eocene (Mángano, Buatois, & Muñiz Guinea, 2005; Strzebonski & Uchman, 2015).

  • GYROPHYLLITES KWASSIZENSIS Glocker, 1841
    Figure 13.1–13.6

  • Material.—KUMIP 314143; two specimens, Cataract Canyon; KUMIP 314162; one specimen, Miners Hollow; KUMIP 314223; one specimen, Antimony Canyon; IBGS PJ-M-022: one specimen, High Creek Canyon, Bear River Range, Utah, USA; IBGS PJ-M-033: one specimen, Miners Hollow.

  • Diagnosis.—Horizontal, straight to club-shaped lobes radiating from single point; may have burrow fills of different color from host lithology, thin ring of disturbed sediment surrounding radiating lobes, and/or lobes that appear bifurcated.

  • Description.—Endorelief, concave epirelief, and convex hyporelief rosettes with 7‐19 straight to club-shaped lobes radiating from a central shaft. Rosettes 16.4–42.5 mm diameter; Lobes 3.9–20.8 mm long, 1.4–11.0 mm wide. Lobes commonly separate but may bifurcate, overlap, or be amalgamated together. Central shaft is only visible on one specimen as a small dark circle, 1.1 mm diameter.

  • Occurrence.—Two lithologies; (1) greenish gray (weathered to brown) siliciclastic silty to sandy shale and may have black to brown dendrites; and (2) dark gray (weather to brown), laminated calcareous silty shale. No visible bedding or laminations present; low to moderate bedding plane disruption (BPBI 2).

  • Associated ichnotaxa.—Planolites beverleyensis, Rusophycus carbonarius, Sagittichnus lincki, and Treptichnus bifurcus.

  • Discussion.—Most specimens assigned to G. kwassizensis occur individually; however, on one slab sample, two endorelief specimens are present and in close proximity (Fig. 13.1–13.2). Only two slab specimens have G. kwassizensis with other ichnotaxa (Fig. 13.3–13.5). Two specimens occur as flat endoreliefs (see Fig. 13.1–13.2). Fill of both rosettes is slightly finer and a lighter color than the surrounding matrix. The lobe shape of each specimen is variable. One specimen has wide lobes with indistinct margins, whereas the other has thinner lobes with distinct margins. At the center of the thin-lobed specimen is a small dark circle, which we interpret as the central tube that would have connected to the next tier and where the tracemaker resided.

  • One collector and donor, Phillip Reese, originally identified a medusoid fossil specimen (Fig. 13.6) as Brooksella Walcott, 1896, which was stored in the KUMIP since 1989 and only recently was reinterpreted as Gyrophyllites by R. A. Robison. Most authors follow the suggestion of Häntzschel (1975) and consider Brooksella to be a body fossil, rather than an ichnofossil. Some authors have retained Brooksella as a valid medusoid ichnogenus (e.g., Willoughby & Robison, 1979; Jensen, 1997), whereas others (e.g., Fürsich & Bromley, 1985) consider Brooksella to be a junior synonym of Dactyloidites Hall, 1886. This specimen has a sandy outer rim surrounding the central, radiating lobes, likely due to the organism having made contact with a sandier layer and stopped excavation (Fürsich & Kennedy, 1975).

  • Willoughby and Robison (1979) reported four specimens of Brooksella from the Spence Shale (Spence Tongue of the Lead Bell Shale of Idaho). Three specimens (Willoughby & Robison, 1979, fig. 1A—C) belong to Gyrophyllites. The fourth (Willoughby & Robison, 1979, fig. 1D) belongs to Dactyloidites as it consists of six radiating lobes with small tubes or tube plugs in the distal ends of the lobes, similar to the specimen of Dactyloidites asterioides Fitch, 1850, figured by Häntzschel (1975, p. 145, fig. 88).

  • Figure 13.

    Gyrophyllites kwassizensis specimens from the Spence Shale. 1–2, Endoreliefs of G. kwassizensis, KUMIP 314143, Cataract Canyon; 1, True color image; 2, False color image; (Arrows indicate central shaft); 3–4, Gyrophyllites kwassizensis preserved in epirelief (3) and hyporelief (4) with Sagittichnus lincki (white arrows) and Treptichnus bifurcus (black arrows), IBGS PJ-M-022, High Creek Canyon, Wasatch Range; 5, Gyrophyllites kwassizensis KUMIP 314223, Antimony Canyon; 6, Gyrophyllites kwassizensis, KUMIP 314162, Miner's Hollow; scale in cm.

    f13_01.jpg

    Ichnogenus HALOPOA Torell, 1870

  • Type ichnospecies.—Halopoa imbricata Torell, 1870, designated by Häntzschel, 1975.

  • Diagnosis.—Long, horizontal burrows covered with irregular longitudinal ridges or wrinkles; may include multiple overlapping cylindrical probes (Uchman, 1998).

  • Discussion.—Halopoa is considered by multiple authors to be very similar to Fucusopsis Palibin in Vasseoevich, 1932, causing each ichnogenus to be transferred back and forth into the other (e.g., Hakes, 1976; Jensen, 1997; Uchman, 1998). Both ichnogenera are described as long, straight to curved, horizontal burrows with longitudinal striations or wrinkles. Hakes (1976) compared and noted that poorly preserved specimens of Fucusopsis, Halopoa, and Scoyenia White, 1929, would be difficult to differentiate. Fucusopsis was synonymized with Palaeophycus and split between P. striatus and P. sulcatus (Pemberton & Frey, 1982). Jensen (1997) argued that Halopoa imbricata was similar to both Fucusopsis and P. sulcatus and partially followed the synonymy of Pemberton and Frey (1982), regarding H. imbricata as a valid ichnospecies within Palaeophycus (i.e., P. imbricatus). Jensen (1997) agreed with the Osgood (1970) interpretation that longitudinal striations on the burrow were caused by sediment deflection and lamination rupture as the tracemaker burrowed through the medium. Jensen (1997) considered them not very useful for ichnotaxonomic assessment as the surficial morphology reflected properties of the sediment. He also noted that there were (rare) spreiten present in H. imbricata but disregarded the fact that spreite are usually an indicator of active burrowing (sensu Fürsich, 1974b).

  • Uchman (1998) argued for the retention of Halopoa, noting the striations of Fucusopsis and Halopoa were likely produced by active digging, passive dragging of body parts due to body shape, or the sediment deflection-lamina rupture method proposed by Osgood (1970). He considered each to represent unique behaviors that generated a unique morphology. Uchman also noted several Halopoa specimens had Teichichnus-like, vertically stacked, overlapping probes (spreiten), but maintained Halopoa and Teichichnus as separate, arguing that Teichichnus generally lacks external ornamentation and that the spreiten in Halopoa were not as developed. He argued against the Pemberton and Frey (1982) synonymy, noting Halopoa lacked any type of wall or lining that would warrant placement within Palaeophycus. We herein follow the Uchman (1998) retention of Halopoa.

  • Three ichnospecies of Halopoa are known; H. annulata (Książkiewicz, 1977), H. imbricata, and H. storeana Uchman, 2001. The primary feature that separates H. imbricata from H. storeana is the orientation of the surficial wrinkles (ridges). The wrinkles of H. imbricata are parallel to subparallel to the trace axis, whereas the wrinkles on H. storeana have a plaited pattern (Uchman, 2001). Halopoa annulata is differentiated from H. imbricata and H. storeana by occasional branching and the presence of transverse annulations producing an undulatory pattern along the burrow (Uchman, 1998, 2001).

  • Commonly interpreted as the feeding burrow of an infaunal deposit feeder, Halopoa tracemakers may include annelid worms, enteropneusts, echiurans, and holothurians (e.g., Hakes, 1976; Uchman, 1998; Zonneveld, Gingras, & Beatty, 2010). Uchman (2001) interpreted Halopoa as a grazing trace. Halopoa is most commonly found in sandstone turbidites of deep-marine flysch deposits (e.g.. Uchman, 1998); however, some have been reported from shallow-marine deposits (e.g., Jensen, 1997) and tidal flats (e.g., Mangano & others, 2002). Halopoa ranges from the early Cambrian to middle Miocene (Jensen, 1997; Uchman, 1998).

  • HALOPOA aff. IMBRICATA Torell, 1870
    Figure 14.1

  • Material.—IBGS LG-M-012: three specimens, Box Elder Canyon, Wellsville Mountains, Utah, USA

  • Diagnosis.—Long, horizontal burrows covered with irregular, longitudinal ridges (Uchman, 1998).

  • Description.—Horizontal, convex hyporelief burrows with irregular-shaped ridges and furrows or wrinkles along the length of the burrow. Burrows 11.0–70.0 mm long, 1.6–2.7 mm wide. Burrows overlap each other to form pseudobranching.

  • Occurrence.—Gray (weathered to brown) calcareous silty shale.

  • Associated ichnotaxa.—Planolites montanus and Treptichnus vagans.

  • Discussion.—Ichnofossils are primarily assigned to Halopoa aff. imbricata due to the longitudinally wrinkled texture of the outer burrow margins and the lack of transverse annulations or a plaited pattern (Fig. 14.1). The nature of the burrow fill is unknown, and the burrows are poorly preserved, convex hyporeliefs. Burrows show overlapping to form pseudobranching, but some possible branching (i.e., secondary successive branching; sensu Keighley & Pickerill, 1995) may be present. Analysis of the fill is needed to confirm the type of branching, if present.

  • Ichnogenus LOCKEIA James, 1879

  • Type ichnospecies.—Lockeia siliquaria, James 1879.

  • Diagnosis.—Amygdaloidal- to ovoid-shaped mounds (convex hyporelief) or depressions (concave epirelief) that taper at one or both ends; surface usually smooth but may be irregular; may have a medial longitudinal crest (hyporelief) or groove (epirelief) (Osgood, 1970; Fillion & Pickerill, 1990; Mángano & others, 2002).

  • Discussion.—Lockeia was the subject of much debate when reintroduced into ichnotaxonomy. Osgood (1970) considered Lockeia as the senior synonym for almond-shaped resting traces and considered Pelecypodichnus Seilacher, 1953a, to be a subjective junior synonym. Numerous authors have followed this suggestion (e.g., Häntzschel, 1975, Hakes 1976, Mángano & others, 2002). Other authors, however, continued to used Pelecypodichnus after Eagar (1974) argued that Lockeia was nomen oblitum citing the International Code of Zoological Nomenclature rule (ICZN 1964, 2nd edition, Article 31) requiring a figure or illustration, alongside the original description, to be a valid taxon (Hakes, 1976; Bromley & Asgaard, 1979; Fillion & Pickerill, 1990). Hakes (1976) retained Lockeia because Article 12 of the ICZN did not require figures for taxa established before 1931 if the original author provided a description or definition. Hakes (1977) later regarded Lockeia as nomen oblitum citing the 50-year rule (ICZN, Article 79) of nonuse of a taxon. Maples and West (1989) noted that Lockeia was used once to validly erect an ichnospecies, Lockeia anticostiana, by Twenhofel (1927) during the supposed 50-year hiatus and, thus, invalidated the argument of Hakes (1977).

  • At least 13 ichnospecies of Lockeia have been proposed: L. amygdaloides (Seilacher, 1953a); L. anticostiana Twenhofel, 1927; L. avalonensis Fillion & Pickerill, 1990; L. cordata Rindsberg, 1994; L. cunctator Schlirf & Uchman in Schlirf, Uchman, & Kümmel, 2001; L. czarnockii (Karaszewski, 1975); L. elongata Yang, 1984; L. gigantus Kim & Kim, 2008; L. hunanensis Zhang & Wang, 1996; L. ornata (Bandel, 1967); L. serialis Seilacher & Seilacher, 1994; L. siliquaria James, 1879; and L. triangulichnus Kim, 1994. Schlirf, Uchman, and Kümmel (2001) and Mángano and others (2002) recently reviewed and compared most of the ichnospecies of Lockeia except for L. gigantus. Lockeia anticostiana was considered as Planolites (Hakes, 1977). Lockeia cunctator is considered a repichnia version of Lockeia similar to and partly as a replacement for L. serialis—considered nomen nudum due to lack of holotype designation and not figured by Seilacher and Seilacher (1994)—but L. cunctator is more similar to Treptichnus bifurcus from where it was transferred by Schlirf, Uchman, and Kümmel (2001)—due to its feather-stitch morphology. Lockeia gigantus was proposed for extremely large, almond-shaped ichnofossils (up to 70 mm long and 30 mm wide) from the Lower Cretaceous lacustrine deposits of Korea (Kim & Kim, 2008).

  • Lockeia is considered a dwelling or resting trace of a bivalve or bivalve-like organism (Seilacher & Seilacher, 1994; Mángano & others, 2002). Lockeia has been reported from shallow marine (e.g., lower delta fronts, and subtidal and intertidal flats), deep marine, and continental lacustrine and fluvial deposits (e.g., Seilacher, 1953a; Hakes, 1976; Bromley & Asgaard, 1979; Crimes & others, 1981; Fillion & Pickerill, 1990; Głuszek, 1995; Hasiotis, 2002, 2004, 2007, 2008; Mángano & others, 2002; Hasiotis & others, 2012). Precambrian Lockeia specimens have been reported (e.g., Narbonne & Aitken, 1990; Jenkins, 1995; McMenamin, 1996), however, Mángano and others (2002) and Jensen, Droser, and Gehling (2006) noted that the traces did not possess the diagnostic characteristics of Lockeia and were likely dubiofossils or body fossils, respectively. The Narbonne and Aitken (1990) specimens show the oval to almond shape characteristic of Lockeia, and, thus, the assignment is justified. Some of the specimens illustrated by McMenamin (1996) may actually be a form of Treptichnus, as they appear to have a feather-stitch morphology characteristic to Treptichnus. Lockeia ranges from the Ediacaran to recent (e.g., Crimes, 1987, 1992; Narbonne & Aitken, 1990; Hasiotis, 2002; Mángano & others, 2002; Jensen, Droser, & Gehling, 2006).

  • Figure 14.

    Halopoa, Monomorphichnus, and Nereites specimens from the Spence Shale. 1, Halopoa aff. imbricata (black arrows) in convex hyporelief, IBGS LG-M-012, Box Elder Canyon; 2, Monomorphichnus bilinearis (arrow) in convex hyporelief with Treptichnus vagans, IBGS PJ-M-031; 3, Monomorphichnus lineatus, convex hyporelief, IBGS PJ-M-012, Miner's Hollow; 4, Monomorphichnus lineatus in concave epirelief, IBGS LG-M-008; 5, Monomorphichnus cf. multilineatus with Cruziana problematica and Lockeia siliquaria convex hyporelief, KUMIP 314228, Miner's Hollow; 6, Nereites cf. macleayi (white arrow) in concave epirelief with Planolites montanus (black arrow), IBGS PJ-M-033, Miner's Hollow; scale: 2 and 6, in mm; 1, 3–5, scale in cm.

    f14_01.jpg

    LOCKEIA SILIQUARIA James, 1879
    Figure 6.5, Figure 19.2

  • Material.—KUMIP 314228; one specimen, Miner's Hollow; IBGS LG-M-003: one specimen; IBGS PJ-M-019; one specimen, Miner's Hollow.

  • Diagnosis.—Amygdaloidal (almond-shaped) convex hyporeliefs, with one or both ends usually tapered to a point, some may be round.

  • Description.—Amygdaloid-shaped (almond-shaped) mound in convex hyporelief; 4.1–6.3 mm long and 2.6–4.4 mm wide.

  • Occurrence.—Two lithologies; (1) Light to dark gray, laminated silty shale with continuous laminations indicating an ii2; and (2) medium to dark gray calcareous, micaceous silty to sandy shale. Bedding planes are highly disrupted with numerous traces indicating BPBI 4–5.

  • Associated ichnotaxa.—Aulichnites isp., Cruziana problematica, Dimorphichnus isp., Monomorphichnus lineatus, M. cf. multilineatus, Phycosiphon incertum, Planolites beverleyensis, P. montanus, Protovirgularia cf. pennatus, Rusophycus carbonarius, Rusophycus cf. cerecedensis, Sagittichnus lincki, and Treptichnus vagans.

  • Discussion.—The Lockeia siliquaria specimen on IBGS PJ-M-019 occurs at the termination of a bilobate concave hyporelief burrow assigned to Aulichnites. The close linear association of the two ichnotaxa suggests they were produced by the same tracemaker. The depth of the L. siliquaria increases toward the opposite side of the intersection of the two ichnotaxa, suggesting the tracemaker produced the Aulichnites and then the Lockeia (see Fig. 6.5). On IBGS LG-M-003, a L. siliquaria specimen is found alongside specimens of Sagittichnus lincki, which are similarly shaped, small ovoid-shaped convex mounds (for discussion see Sagittichnus, p. 36). The two morphologies can be distinguished by size, as L. siliquaria is larger than the Sagittichnus lincki specimens.

  • Ichnogenus MONOMORPHICHNUS Crimes, 1970b

  • Type ichnospecies.—Monomorphichnus bilinearis Crimes, 1970b.

  • Diagnosis.—Series of straight to sigmoidal, parallel or intersecting, laterally repeating striations in isolated or grouped sets; typically preserved in convex hyporelief (Crimes, 1970b; Fillion & Pickerill, 1990; Keighley & Pickerill, 1998).

  • Discussion.—Crimes (1970b) established Monomorphichnus for surficial striations produced by bottom-current-propelled trilobites raking the sediment surface with their endopodite claws. He noted that these striations were similar to Dimorphichnus but lacked the characteristic blunt impressions (Crimes, 1970b; Fillion & Pickerill, 1990; Jensen, 1997). Monomorphichnus maybe a junior synonym of Ctenichnites Matthew, 1891, Eoichnites Matthew, 1891, Medusichnites Matthew, 1891, or Taonichnites Matthew in Selwyn, 1890; however, their ichnotaxonomic status is unclear as some authors have considered them dubiofossils or pseudofossils, whereas others considered them valid ichnotaxa (see Fillion & Pickerill, 1990, for full discussion).

  • Since Monomorphichnus was established at least 15 ichnospecies have been proposed and differentiated by the number of striations present: M. bilinearis Crimes, 1970b; M. biserialis Mikuláš, 1995; M. cretacea Badve & Ghare, 1980; M. devonicus Yang & Hu in Yang, Hu, & Sun, 1987; M. gaopoensis Yang, Yin, & He, 1982; M. gregarius, Pandey & others, 2014; M. henanensis Yang & Wang, 1991; M. intersectus Fillion & Pickerill, 1990; M. lineatus Crime & others, 1977; M. monolinearis Shah & Sudan, 1983; M. multilineatus Alpert, 1976a; M. pectenensis Legg, 1985; M. podolicus Uchman & others, 2004; M. semilineatus Mikuláš, 1995; and M. sinus Gibb, Chatterton, & Pemberton, 2009. Monomorphichnus cretacea and M. gaopoensis are considered inorganic tool marks or a combination of organic and inorganic structures (e.g., Fillion & Pickerill, 1990; Uchman & others, 2004). Fillion and Pickerill (1990) found M. monolinearis to be a junior synonym of M. lineatus. Monomorphichnus podolicus was synonymized with Cruziana omanica Seilacher, 1970, due to its tendency to occur bilobate (Gibb, Chatterton, & Pemberton, 2009). Monomorphichnus gregarius was introduced by Pandey and others (2014) for highly overlapping sets of 4 striations; however, the holotype has sets of 4–6 striations, most of which occur in sets of 6, with central striations being more prominent, and crosscut other sets, which suggests affinities to both M. multilineatus and M. intersectus. We, therefore, regard M. gregarius and M. intersectus as subjective junior synonyms of M. multilineatus.

  • Monomorphichnus is considered a locomotion or grazing trace (Crimes, 1970b; Crimes & others, 1977), and often has been attributed to trilobites (e.g., Crimes, 1970b; Alpert, 1976a), but other arthropods (e.g., eurypterids and xiphosurids) have also been proposed as possible tracemakers (Romano & Meléndez, 1985; Jensen, 1997). Osgood (1970) suggested the grazing interpretation was an inefficient feeding strategy and that Monomorphichnus was likely produced by an arthropod trying to stabilize itself while caught in turbulent bottom-water currents. A recent neoichnological study by Jones (2016) has shown several Monomorphichnus-like traces produced by bats via ground-based locomotive and searching behaviors. Monomorphichnus has been reported from shallow- and deep-marine, and continental deposits (e.g., Crimes, 1970b; Crimes & others, 1977; Keighley & Pickerill, 1998). The earliest occurrence of Monomorphichnus has been thought to be in units previously referred to as Vendian—now known as the Ediacaran—by Crimes (1987, 1992). Reports of Monomorphichnus from latest Neoproterozoic strata by Jenkins (1995) and Waggoner and Hagadorn (2002) were reinterpreted by Jenson, Droser, and Gehling (2006) as Radulichnus and tool marks or a trace fossil(?), respectively. However, the redefinition of the Ediacaran (Neoproterozoic)-Cambrian (Paleozoic) boundary based on the occurrence of Treptichnus pedum may define the range of Monomorphichnus as Cambrian to recent (e.g., Jensen, 1997; MacNaughton & Narbonne, 1999; Jenson, Droser, & Gehling, 2006; Landing & others, 2007; Hasiotis, 2012).

  • MONOMORPHICHNUS BILINEARIS Crimes, 1970b
    Figure 14.2, Figure 23.5

  • Material.—IBGS PJ-M-031; five specimens, Miners Hollow.

  • Diagnosis.—Pairs of parallel, straight to slightly sigmoidal striations with one striation more prominent than the other, and sometimes repeated laterally (Crimes 1970b; Fillion & Pickerill, 1990).

  • Description.—Paired sigmoidal striations in convex hyporelief. Striations 12.1–45.2 mm long, 0.6–1.7 mm wide, and spaced 1.0–1.3 mm apart.

  • Occurrence.—Gray (weathered to brown), micaceous silty shale.

  • Associated ichnotaxa.—Treptichnus vagans.

  • Discussion.—Monomorphichnus bilinearis are only present on IBGS PJ-M-031 alongside Treptichnus vagans. The striations were assigned to this ichnogenus due to their sigmoidal shape and tendency to occur in pairs. Some M. bilinearis specimens are cross cut by Treptichnus vagans specimens (Fig. 14.2), thus, indicating the striations were produced first, followed by the construction of the Treptichnus vagans.

  • MONOMORPHICHNUS LINEATUS Crimes, & others, 1977
    Figure 14.3–14.4

  • Material.—KUMIP 314228; one specimen, Miners Hollow; IBGS LG-M-008 and LG-M-009 (part and counterpart): one specimen, Spence Shale; IBGS PJ-M-012: one specimen, Miners Hollow.

  • Diagnosis.—Individual, straight to slightly sigmoidal striations that can be repeated laterally (Crimes & others, 1977; Fillion & Pickerill, 1990).

  • Description.—Sigmoidal to slightly curved striations, some may be bifid, in convex hyporelief and concave epirelief. Striations 5.4– 36.1 mm long, 0.5–2.1 mm wide. One row of repeated striations is 50.6 mm long, 9.7 mm wide, and spaced 1.5–1.9 mm apart. Striations may have blunt ends and sharply taper on the other.

  • Occurrence.—Two lithologies: (1) greenish gray (weathered to tan or brown), micaceous silty shale; and (2) gray, silty shale with laminations of light and dark gray, silty to sandy, siliciclastic to carbonate shale.

  • Associated ichnotaxa.—Cruziana problematica, Lockeia siliquaria, Monomorphichnus cf. multilineatus, Planolites beverleyensis, P. montanus, Protovirgularia dichotoma, Rusophycus carbonarius, Rusophycus cf. cerecedensis, Treptichnus bifurcus, and T. vagans.

  • Discussion.—The specimen on IBGS PJ-M-012 is differentiated from Cruziana billingsi Fillion & Pickerill, 1990, by the arrangement of striations in a single track row—whereas C. billingsi is bilobate—and is almost identical to the holotype illustrated by Crimes and others (1977, p. 107, pl. 3b) (Fig. 14.3). Specimens on IBGS LG-M-008 and LG-M-009 occur with no other traces (Fig. 14.4).

  • MONOMORPHICHNUS cf. MULTILINEATUS Alpert, 1976a
    Figure 14.5

  • Material.—KUMIP 314228: one specimen, Miner's Hollow.

  • Diagnosis.—Parallel, straight to sigmoidal striations grouped in sets of 5 to 6, with deeper and thicker striations in center of group (Alpert, 1976a).

  • Description.—Horizontal, sigmoidal striations (convex hyporelief) grouped in sets of 2–4, spaced 1.6–1.8 mm apart with one striation more prominent than the others. Striations 3.0–14.2 mm long, 0.3–0.7 mm wide, and spaced 0.5–1.1 mm apart.

  • Occurrence.—Greenish gray (weathered to tan or brown), micaceous silty shale.

  • Associated ichnotaxa.—Cruziana problematica, Lockeia siliquaria, Monomorphichnus lineatus, Planolites beverleyensis, P. montanus, Rusophycus carbonarius, R. cf. cerecedensis, and Treptichnus bifurcus.

  • Discussion.—The specimen has striations grouped in pairs assignable to M. bilinearis, but others grouped in triplets and quadruplets, which are assignable to M. multilineatus. A specimen illustrated by Fillion and Pickerill (1990, pl. 10, fig. 3) has several bundles of 2–3 striations mixed with the typical 4–6 striation bundles. The similarity between the Spence Shale specimen and the Fillion and Pickerill (1990) specimen justifies assignment to M. multilineatus. Another Monomorphichnus ichnospecies that the Spence Shale specimen resembles is M. semilineatus Mikuláš, 1995, which is characterized as curved to straight sigmoidal striations in groups of 2–10 (Mikuláš, 1995, pl. 1 & 3, fig. 1C). Monomorphichnus semilineatus, however, appears to be morphologically variable with bundle sets that are indistinguishable from other Monomorphichnus ichnospecies. We, therefore, consider M. semilineatus to be an amalgam of several Monomorphichnus ichnospecies and no valid use to ichnotaxonomy. Monomorphichnus multilineatus also resembles the coarse striations of Rusophycus dispar Linnarsson, 1869, like those figured by Jensen (1990, fig. 1) (Fig. 14.5); however, the specimen lacks the bidirectionality and bilobate shape typical of R. dispar.

  • Ichnogenus NEREITES MacLeay in Murchison, 1839

  • Type ichnospecies.—Nereites cambrensis MacLeay 1839 in Murchison (1839, p. 700).

  • Diagnosis.—Curved, winding to regularly meandering, unbranched, horizontal trails, with a medial backfilled tunnel flanked by an even to lobate zone of reworked sediment (Uchman, 1995; Mángano & others, 2000, 2002).

  • Discussion.—A long-lasting debate in ichnotaxonomy has been raging regarding status of the ichnotaxa Nereites MacLeay, 1839 in Murchison, 1839; Neonereites Seilacher, 1960; and Scalarituba Weller, 1899. Numerous authors have suggested that Nereites is the senior synonym of Neonereites and Scalarituba, arguing that both are preservational variants of Nereites (e.g., Chamberlain, 1971; Chamberlain & Clark, 1973; D'Alessandro & Bromley, 1987; Devera, 1989; Rindsberg, 1994; Uchman 1995; Mángano & others, 2000, 2002). Some authors, however, retain or advocate for the retention of Neonereites as a separate ichnotaxon (e.g., Benton, 1982; Fillion & Pickerill, 1990; Pickerill, 1991).

  • Though long suggested, Uchman (1995) was one of the few to formally place Neonereites and Scalarituba within Nereites. He also suggested that the three Neonereites ichnospecies (N. biserialis Seilacher, 1960; N. multiserialis Pickerill & Harland, 1988; and N. uniserialis Seliacher, 1960) should be used informally as subichnospecies to describe associated preservational variation. Helminthoida Schafhäutl, 1851, was also synonymized under Nereites because Uchman (1995) noted Nereites-like marginal lobes in the type specimen. Seilacher (1962) suggested that Helminthoida and Neonereites were related with Neonereites being a preservational variant of Helminthoida in sand-rich environments. We, however, suggest retaining Helminthoida due to its high-sinuosity, tightly meandering, and repetitive pattern in morphology that is distinctive and diagnostic of this ichnotaxon, which is morphologically related to Helminthopsis Heer, 1877.

  • Nereites is interpreted as a deposit-feeding or grazing trace (e.g., Uchman, 1995; Mángano & others, 2000, 2002). Commonly proposed tracemakers include annelid, enteropneust, and polychaete worms (e.g., Seilacher, 1960; Rindsberg, 1994; Uchman 1995; Mángano & others, 2000, 2002); however, gastropods, arthropods, and echinoderms (e.g., holothurians) have also been proposed (e.g., Rindsberg, 1994). Although the namesake of the deep marine Nereites Ichnofacies, Nereites has been reported from shallow-marine (e.g., lagoon, shoreface, tidal flats) and deep-marine settings (e.g., flysch) (e.g., Hakes, 1976; McCann & Pickerill, 1988; Uchman, 1995, 1998; Mángano & others, 2000). Nereites is common in both shallow- and deep-marine Paleozoic deposits but became almost exclusively deep marine in Mesozoic and Cenozoic deposits (e.g., Uchman, 1995; Mángano & others, 2000). Mángano and others (2002) argued that the lacustrine Nereites specimens figured by Hu, Wang, and Goldring (1998) did not fit the diagnostic criteria for Nereites and belong in Vagorichnus Buatois and others, 1995. We, however, consider Vagorichnus to be a junior synonym of Walpia White, 1929, based on morphologic similarities. We suggest that the specimens figured by Hu, Wang, and Goldring (1998) have morphologic features assignable to Walpia, which are typical of burrows produced by modern mud-loving beetles and some spiders just above the sediment-water interface (Hasiotis, 2002, 2004, 2008). Nereites has been reported from the Vendian (i.e., Edicaran) (e.g., Crimes & Germs, 1982; Jenkins, 1995); however, Jensen, Droser, and Gehling (2006) considered those specimens to be a form of Archaeonassa. Yet the photograph of the Nereites specimen in Crimes and Germs (1982) does show a central furrow flanked by ridges that are subtly lobate that grade into strongly hemispherical lobes typical of several Nereites ichnospecies; thus, we consider this specimen to be Nereites. The stratigraphic position of this specimen, however, is in the Vingerbreek Member of the Nudaus Formation of the lower part of the Schwarzrand Subgroup, which is Ediacaran in age, based on the co-occurrence of body fossils (e.g., Cohen & others, 2009). Nereites, therefore, ranges from the Ediacaran to recent (e.g., Crimes & Germs, 1982; Crimes, 1992; Mángano & others, 2000; Uchman, 1995).

  • NEREITES cf. MACLEAYI MacLeay in Murchison, 1839
    Figure 14.6

  • Material.—IBGS PJ-M-033: one specimen (part and counterpart), Miner's Hollow.

  • Diagnosis.—Small, straight to meandering, concave furrow (epirelief) or convex burrow (hyporelief) flanked by small, semicircular lobes along furrow margin (McCann & Pickerill, 1988).

  • Description.—Straight, concave furrow flanked by small, semicircular lobes 24.1 mm long, 2.5–4.5 mm wide. Furrow 1.1–3.0 mm wide, and lobes 1.4–1.8 mm wide (from furrow margin). Furrow has a serial, spherical-chambered expression, chamber diameter 1.4–3.0 mm.

  • Occurrence.—Gray (weathered to brown), siliciclastic silty to sandy shale.

  • Associated ichnotaxa.—Archaeonassa fossulata, Gyrophyllites kwassizensis, and Planolites montanus.

  • Discussion.—The assignment to Nereites cf. macleayi was based primarily on the presence of small, semicircular lobes present along the furrow margin (Fig. 14.6). The furrow also has a serialchamberlike appearance similar to Neonereites uniserialis; however, since Neonereites was synonymized under Nereites, assignment to Neonereites is untenable. Assignment to Nereites missouriensis may be justified by the presence of the serial chambers; yet, no meniscate backfill typical of N. missouriensis is observed in the specimen.

  • Ichnogenus PHYCODES Richter, 1850

  • Type ichnospecies.—Phycodes circinatus Richter, 1853.

  • Diagnosis.—Horizontal to subhorizontal, cylindrical to U-shaped burrows with dichotomously branched tunnels forming bundles (Fillion & Pickerill, 1990; Knaust, 2007).

  • Discussion.—Since Richter (1850) originally designated Phycodes for bundled structures regarded as fucoids, Phycodes has undergone several revisions to its present-day status as an ichnofossil (see Fillion & Pickerill, 1990; Han & Pickerill, 1994b; Jensen, 1997). Phycodes has been interpreted as a deposit-feeding trace of annelid worms (Fillion & Pickerill, 1990). Phycodes has been considered to be a good indicator for shallow-marine settings and indicative of the Cruziana Ichnofacies, but Phycodes has been reported from brackish and deep-water deposits as well (e.g., Hakes, 1985; Fillion & Pickerill, 1990; Han & Pickerill, 1994a; Jackson, Hasiotis, & Flaig, 2016). Phycodes ranges from the early Cambrian to the Miocene (Crimes, 1987, 1992; Han & Pickerill, 1994a).

  • Figure 15.

    Phycodes and Planolites specimens from the Spence Shale. 1, Phycodes curvipalmatum, in partial convex hyporelief and endorelief, IBGS PJ-M-005, Miner's Hollow; 2, Planolites annularis in concave epirelief, IBGS PJ-M-001, Miner's Hollow; 3–5, Planolites beverleyensis: 3, Convex hyporelief, IBGS LG-M-005; 4, Concave hyporelief, IBGS LG-M-001; 5, Self-crossing specimen (white arrows) with Cruziana problematica (black arrow) and Rusophycus cf. cerecedensis (circle) in convex hyporelief, KUMIP 315228, Miner's Hollow; 6, Planolites montanus in convex hyporelief, IBGS LG-M-012, Box Elder Canyon; scale in cm.

    f15_01.jpg

    PHYCODES CURVI PALMATUM Pollard, 1981
    Figure 15.1

  • Material.—IBGS PJ-M-005: one specimen, Miner's Hollow float.

  • Diagnosis.—Thin, short, rounded, horizontal palmate or digitate burrows that originate from the same point (Pollard, 1981).

  • Description.—Small trifid-branched system of short burrows in convex epirelief and partial endorelief. Burrows range from 4.6–16.9 mm long and 2.0–2.5 mm wide. Burrow fill of two branches are exposed with burrow walls 0.3–0.8 mm thick and is similar to the host lithology.

  • Occurrence.—Tan to light brown, siliciclastic silty shale.

  • Associated ichnotaxa.—Archaeonassa jamisoni and Taenidium cf. satanassi.

  • Discussion.—The specimen consists of a triplet of burrows exhibiting primary successive branching (sensu D'Alessandro & Bromley, 1987; Keighley & Pickerill, 1995). (Fig. 15.1). The specimen is assigned to P. curvipalmatum and not P. palmatus because the burrows are both short and narrow, whereas P. palmatus Hall, 1852 has long and wide burrows. Two branches are preserved mostly as endoreliefs with the burrow walls as the only significant structures remaining, whereas one branch is in convex epirelief.

  • Ichnogenus PHYCOSIPHON Fischer-Ooster, 1858

  • Type ichnospecies.—Phycosiphon incertum Fischer-Ooster, 1858, by original monotypy.

  • Diagnosis.—Small, oblique or parallel to bedding, spreiten-filled burrow systems comprised of protrusive U-shaped lobes with dark, finer grained cores and light, coarser grained mantles; lobes may be nearly vertical to bedding; spreiten may not be visible (Wetzel & Bromley, 1994; Głuszek, 1998; Uchman, 1998).

  • Discussion.—Like most ichnofossils, Phycosiphon was originally interpreted as fossilized algae. More recently, however, it was interpreted as a complex burrow system of a deposit feeder, typically in dysoxic sediments (e.g., Ekdale & Mason, 1988; Uchman, 1998; Naruse & Nifuku, 2008). Wetzel and Bromley (1994) noted two general lobe arrangements occur in Phycosiphon, influenced by the host lithology; (1) lobes are parallel or subparallel to bedding in laminated sands and silts (exaggerated by compaction); and (2) lobes are randomly to vertically oriented in muddy and homogenous sediments. Wetzel and Bromley (1994) also compared Phycosiphon to Anconichnus Kern, 1978, because both are mantled, spreiten-filled, U-shaped burrow systems, and they decided that Anconichnus was a junior synonym of Phycosiphon.

  • Phycosiphon was monotypic with P. incertum as its sole ichnospecies until Uchman (1998) synonymized Muensteria hamata Fischer-Ooster, 1858, under Phycosiphon as P. hamata, and later joined by Muensteria geniculata Sternberg, 1833 by Uchman (1999) as P. geniculatum. Phycosiphon hamata differs from P. incertum with its more regularly shaped lobes, larger size, and J- to U-shaped lobes. Uchman (1998) also warned that P. hamata should not confused with Zoophycos, which occur in multiple levels, whereas P. hamata occurs on only one. Phycosiphon geniculatum differs from P. hamata and P. incertum by having radially arranged lobes with one margin well defined, usually concave, and the other margin is convex and highly lob ate and indistinct. Naruse and Nifuku (2008) demonstrated that the elliptical burrow cross-sections of Phycosiphon could be used to determine the paleoslope inclination of a deposit.

  • Phycosiphon is interpreted a trace of a deposit-feeding, wormlike organism (Wetzel & Bromley, 1994). Phycosiphon occurs in continental-shelf slopes, submarine fans, turbidites, and flysch deposits (e.g., Uchman, 1998; Naruse & Nifuku, 2008; Rajchel & Uchman, 2012). Recent studies have found that Phycosiphon tracemakers are early colonizers of the upper portions of turbidite deposits when bottom waters are fully oxygenated (e.g., Wetzel & Uchman, 2001; Naruse & Nifuku, 2008). Phycosiphon ranges from the early Cambrian to recent (Fu, 1991; Naruse & Nifuku, 2008).

  • PHYCOSIPHON INCERTUM Fischer-Ooster, 1858
    Figure 16.1–16.6

  • Material.—IBGS LG-M-007; seven specimens; IBGS PJ-M-019; five specimen, Miner's Hollow.

  • Diagnosis.—Small, oblique or parallel to bedding, spreiten-filled burrow systems comprised of protrusive U-shaped lobes with dark, fine-grained cores and light, coarse-grained mantles (Wetzel & Bromley, 1994; Głuszek, 1998; Uchman, 1998).

  • Description.—Mantled, endorelief burrows with elliptical to U-shaped cross sections. Light gray mantles 0.1–0.4 mm thick, average thickness 0.2 mm. Dark gray cores 0.3–1.0 mm thick, 1.1–7.8 mm wide.

  • Occurrence.—Two lithologies; (1) green, fine-grained siliciclastic sandstone, weathered to tan; and (2) laminated mudstone of alternating light-gray and dark-gray laminations. Laminations on IBGS PJ-M-019 are continuous with very little bioturbation occurring to disrupt them (ii2), but on IBGS LG-M-007, the laminations are moderately disrupted (ii3–4).

  • Associated ichnotaxa.—Aulichnites isp., Dimorphichnus isp., Lockeia siliquaria, Protovirgularia cf. pennatus, and Treptichnus vagans.

  • Discussion.—Phycosiphon incertum present on IBGS LG-M-007 in cross section show a light gray to white mantle and some spreite within the burrow fill (Fig. 16.1–16.4). Some spreiten are visible in longitudinal cross section (Fig. 16.1–16.2), but are most visible in specimens with transverse cross sections (Fig. 16.3–16.4). The sediment of IBGS LG-M-007 is mostly pale green to white finegrained sandstone, while the burrow fill is composed of fine- to very fine-grained, gray to black sandstone. Sample IBGS PJ-M-019 has several specimens of P. incertum on the cut side of the samples (Fig. 16.5–16.6). The mantle surrounding some of the IBGS PJ-M-019 burrows is not very noticeable, possibly due to their small size and compaction.

  • Ichnogenus PLANOLITES Nicholson, 1873

  • Type ichnospecies.—Planolites beverleyensis Billings, 1862 (=Planolites vulgaris Nicholson & Hinde, 1875, junior synonym, Pemberton & Frey, 1982).

  • Diagnosis.—Unlined to rarely lined, rarely branching, straight to tortuous burrows with smooth to irregular walls and circular to elliptical cross sections; infill unstructured and may differ from host-rock lithology (Pemberton & Frey, 1982; Fillion & Pickerill, 1990; Uchman, 1998).

  • Discussion.—Ichnotaxonomy still has numerous problems with differentiating between certain ichnotaxa, including Palaeophycus Hall, 1847, and Planolites (e.g., Osgood, 1970; Häntzschel, 1975; Pemberton & Frey, 1982). In an attempt to resolve those problems, Pemberton and Frey (1982) reexamined both ichnogenera and established standard diagnostic criteria for differentiating them: (1) burrows lack wall linings; and (2) burrows have different color and texture from host-rock lithology that indicate active infilling. Another criterion suggested to help identify Planolites is the lack of systematic branching or enlargements around branch sites (Fillion & Pickerill, 1990). Keighley and Pickerill (1995) argued against the use of active vs. passive infill and suggested that the presence or lack of a wall lining should be the primary diagnostic criterion for Palaeophycus and Planolites, respectively. Keighley and Pickerill (1997) also recommended synonymizing P. montanus under P. beverleyensis and argued that the size criterion used to separate the two ichnospecies was invalid; however, most authors have ignored the Keighley and Pickerill (1997) recommendation and continue to use both P. beverleyensis and P. montanus (e.g., Pickerill & Fyffe, 1999; Uchman, 1999; Hofmann & others, 2012).

  • Planolites is typically interpreted as the trace of a deposit-feeding marine or freshwater worm (e.g., Häntzschel, 1975; Fillion & Pickerill, 1990); however, soil arthropods and worms have been suggested as possible tracemakers in continental deposits (e.g., Ekdale, Bromley, & Loope, 2007; Hasiotis, 2004, 2008; Smith & others, 2008a, 2009). Planolites is a facies-crossing ichnogenus and has been report from shallow- to deep-marine and continental deposits (e.g., alluvial, fluvial, lacustrine, and eolian) (e.g., Chamberlain, 1971, 1975, 1977; Fillion & Pickerill, 1990; Keighley & Pickerill, 1997; Kim & others, 2005; Ekdale, Bromley, & Loope, 2007; Bohacs, Hasiotis, & Demko, 2007; Hembree & Hasiotis, 2007; Hofmann & others, 2012). Planolites ranges from the Ediacaran to recent (Häntzschel, 1975; Crimes, 1987, 1992; Uchman, 1998).

  • Figure 16.

    Phycosiphon incertum specimens from the Spence Shale. 1–4, Full relief and cross sections of P. incertum, IBGS LG-M-007; 5–6, Cross sections of P. incertum, IBGS PJ-M-019, Miner's Hollow; scale bars in mm.

    f16_01.jpg

    PLANOLITES ANNULARIUS Walcott, 1890
    Figure 15.2

  • Material.—KUMIP 314229: two specimens, Miner's Hollow; IBGS PJ-M-001: four specimens, Miner's Hollow.

  • Diagnosis.—Horizontal, straight to curved, subcylindrical burrows with pronounced annulations (Pemberton & Frey, 1982; Fillion & Pickerill, 1990).

  • Description.—Simple, straight to curved burrows in concave epirelief or convex hyporelief with transverse constrictions forming numerous short chambers (1.1–2.0 mm long). Burrows 17.1–125.2 mm long, 0.8–1.5 mm wide. On IBGS PJ-M-001, a reddish brown halo is present along some burrows and extends 0.9–2.3 mm from burrow margin.

  • Occurrence.—Gray to dark gray, massive siliciclastic shale.

  • Associated ichnotaxa.—Bergaueria hemispherica, Cruziana barbata, and Rusophycus carbonarius.

  • Discussion.—The burrows were formed by peristaltic movement of a wormlike tracemaker that resulted in the serial-chambered expression (Pemberton & Frey, 1982). Most P. annularius specimens have a reddish brown oxidation halo surrounding the burrow, indicating poorly oxygenated sediments near the time of construction (e.g., Ekdale, Bromley, & Pemberton, 1984; Bromley, 1996; Forster, 1996) (see Fig. 15.2). One burrow has the reddish brown halo for about half its length and the entire width extending to ∼3 mm from the burrow center, but also loses the annulated chambers where the halo is present. The change from an annulated burrow with or without a halo to a fully haloed, smooth burrow could be a transition from P. annularius to P. montanus representing a change in oxygen and nutrient availability in the sediment (e.g., Pemberton & Frey, 1982; Forster, 1996).

  • PLANOLITES BEVERLEYENSIS (Billings, 1862)
    Figure 15.3–15.5

  • Material.—KUMIP 314223: one specimen, Antimony Canyon; KUMIP 314228: one specimen, Miner's Hollow;s IBGS LG-M-001: one specimen; IBGS LG-M-005: four specimens.

  • Diagnosis.—Large, smooth, straight to gently curved or undulated cylindrical burrows with unstructured backfill and lacking wall linings (Pemberton & Frey, 1982; Fillion & Pickerill, 1990; Keighley & Pickerill, 1995).

  • Description.—Convex hyporelief or concave epirelief, straight to gently curved cylindrical burrow; however, some are contorted and overlap. Burrows 10.3–50.4 mm long and 3.0–6.3 mm wide. No wall lining is visible.

  • Occurrence.—Light to dark gray (weathered to light brown or tan), siliciclastic silty shale.

  • Associated ichnotaxa.—Cruziana problematica, Gyrophyllites kwassizensis, Lockeia siliquaria, Monomorphichnus lineatus, M. cf. multilineatus, Planolites montanus, Rusophycus carbonarius, and Treptichnus bifurcus.

  • Discussion.—Planolites beverleyensis is typically differentiated from others ichnospecies of Planolites by its larger burrow diameter (> 5 mm), its generally straighter course, and a lack of annulations (Billings, 1862; Pemberton & Frey, 1982). Though most burrow diameters are < 5 mm, specimens assigned to P. beverleyensis are significantly larger and straighter than any specimen assigned to P. montanus. One specimen of P. beverleyensis appears to record a predation-prey interaction with a Rusophycus cf. cerecedensis (Fig. 15.5) (for discussion see Rusophycus cf. cerecedensis p. 34).

  • PLANOLITES MONTANUS Richter, 1937
    Figures 10.3, 14.6, 15.6, 18.3, 18.6, 22.3

  • Material.—KUMIP 204523 A and B: four specimens, Miner's Hollow; KUMIP 314122: one specimen, Antimony Canyon; KUMIP 314222 B: 13 specimens, Miner's Hollow; KUMIP 314228: 11 specimens, Miner's Hollow; IBGS LG-M-010: four specimens; IBGS LG-M-011: four specimens; IBGS LG-M-012: two specimens; IBGS LG-M-013: three specimens; IBGS PJ-M-001: one specimen, Miner's Hollow; IBGS PJ-M-004: two specimens, Miner's Hollow; IBGS PJ-M-005: five specimens, Miner's Hollow float; IBGS PJ-M-007: six specimens, Miner's Hollow; IBGS PJ-M-010: three specimens, Miner's Hollow IBGS PJ-M-011: three specimens, Spence Tongue of the Lead Bell Shale, Oneida Narrows, Idaho; IBGS PJ-M-013: one specimen, Miner's Hollow; IBGS PJ-M-014: one specimen, Miner's Hollow; IBGS PJ-M-016: two specimens, Miner's Hollow; IBGS PJ-M-020: one specimen, Miner's Hollow; IBGS PJ-M-023: two specimens, Miner's Hollow; IBGS PJ-M-024: one specimen, Miner's Hollow; IBGS PJ-M-027: two specimens, Miner's Hollow; IBGS PJ-M-030: one specimen, Miner's Hollow; IBGS PJ-M-033: two specimens.

  • Diagnosis.—Relatively small, curved to tortuous, cylindrical to subcylindrical burrows lacking wall linings (Pemberton & Frey, 1982; Fillion & Pickerill, 1990; Keighley & Pickerill, 1995).

  • Description.—Small, smooth burrows that are generally straight but sometimes sharply bent, curved, or contorted. Burrows 12.5–73.2 mm long and 0.7–5.4 mm wide.

  • Occurrence.—Laminated light to dark gray or dark gray (weathered to tan or brown) to pale greenish gray, calcareous or siliciclastic silty shale.

  • Associated ichnotaxa.—Archaeonassa fossulata, Cruziana problematica, Halopoa aff. imbricata, Lockeia siliquaria, Monomorphichnus lineatus, M. cf. multilineatus, Nereites cf. macleayi, Planolites beverleyensis, Rusophycus carbonarius, R. cf. cerecedensis, Treptichnus bifurcus, and Treptichnus vagans.

  • Discussion.—Planolites montanus is one of the most common ichnofossils from the Spence Shale. Burrows assigned to this ichnotaxon exhibit simple, smooth margins, fill different from the host rock, a lack of distinct walls, and diameters typically < 5 mm. Some specimens of P. montanus occur in a Treptichnus-like morphology of short burrows that alternate very loosely around a central axis.

  • Ichnogenus PROTOVIRGULARIA M‘Coy, 1850

  • Type ichnospecies.—Protovirgularia dichotoma M‘Coy, 1850.

  • Diagnosis.—Unbranched, straight to slightly curved trails with medial ridge or furrow and paired, wedge-shaped, lateral projections from ridge or furrow (Han & Pickerill, 1994b).

  • Discussion.—Protovirgularia was originally interpreted by M‘Coy (1850) as a body fossil of octocoral and graptolites. Han and Pickerill (1994b), Seilacher and Seilacher (1994), and Uchman (1998) have reviewed Protovirgularia. Han and Pickerill (1994b) reassessed the four previously established ichnospecies of Protovirgularia, and found that: (1) P. dichotoma M‘Coy, 1850 was the only valid ichnospecies; (2) P. mongraensis Chiplonkar & Badve, 1970, and P. nereitarum (Richter, 1871) as junior synonyms of P. dichotoma; and (3) P. harknessi Lapworth, 1870, was nomen nudum and not valid because it was not described or figured when originally proposed. Seilacher and Seilacher (1994) demonstrated via neoichnological experiments that protobranch bivalves and scaphopods were the primary producers of Protovirgularia and also synonymized Imbrichnus Hallam, 1970, Pennatulites De Stefani, 1885, Uchirites Macsotay, 1967, and Walcottia Miller & Dyer, 1878, under Protovirgularia. Uchman (1998) also expanded Protovirgularia to include some ichnospecies of Gyrochorte, Nereites, Rhabdoglyphus Vassoevich, 1951, and Tuberculichnus Książkiewicz, 1977, as junior synonyms.

  • Protovirgularia is interpreted as a push-pull locomotion and feeding trace of bivalves and scaphopods (e.g., Han & Pickerill, 1994b; Seilacher & Seilacher, 1994). Protovirgularia has been reported from shallow marine, deep marine (e.g., turbidites), and brackish water deposits (e.g., deltas, estuaries, and tidal flats) (e.g., Han & Pickerill, 1994b; Seilacher & Seilacher, 1994; Carmona & others, 2010; Jackson, Hasiotis, & Flaig, 2016). Protovirgularia is sometimes suggestive of salinity, sedimentation rate, and turbidity fluctuations as well as possible oxygen depletion (Carmona & others, 2010). Protovirgularia ranges from the early Cambrian to recent (Seilacher & Seilacher, 1994; Orłowski & Zylińska, 2002).

  • PROTOVIRGULARIA DICHOTOMA M‘Coy, 1850
    Figure 17.1

  • Material.—KUMIP 314233: one specimen.

  • Diagnosis.—Straight, bilobate trails with medial furrow and paired, convex, chevronlike, wedge-shaped projections oblique from furrow.

  • Description.—Specimen 29.1 mm long and 4.1 mm wide. The chevronlike, wedge-shaped projections range 2.7–4.2 mm long and 1.5–2.3 mm wide. Projection sets have a 45–55° V-shaped angle. Faint striations are present on projections.

  • Occurrence.—Dark gray (weathered to tan), calcareous silty shale with very thin siliciclastic mud with possible swaley crossstratification.

  • Associated ichnotaxa.—Monomorphichnus lineatus and Treptichnus vagans.

  • Discussion.—Though similar to Didymaulichnus Young, 1972, due to its bilobate shape and seemingly smooth projections, the P. dichotoma specimen has a chevronlike morphology most similar to Protovirgularia morphologic variant 5 of Carmona and others (2010, fig. 3.8 & 4) and experimental undertraces analogous to P. dichotoma illustrated by Seilacher and Seilacher (1994, pl. 1, fig. a) (Fig. 17.1). Protovirgularia specimens illustrated by Fernández, Pazos, and Aguirre-Urreta (2010) show intergradation between P. dichotoma and P. rugosa, and the lateral projections of the Spence Shale P. dichotoma are more oblique and wedgelike, and thus are more similar to P. dichotoma. However, the morphologic characteristic that separates P. dichotoma and P. rugosa is the presence of a Lockeia-like object at the termination of P. rugosa (sensu Seilacher & Seilacher, 1994; Uchman, 1998), which the Spence Shale P. dichotoma specimen lacks.

  • PROTOVIRGULARIA cf. PENNATUS (Eichwald, 1860)
    Figure 6.5, Figure 17.2–17.4

  • Material.—KUMIP 204521 A and B: one specimen; IBGS PJ-M-019: Miner's Hollow.

  • Diagnosis.—Straight to winding, bilobate, chevronlike ribbon trace with medial ridge.

  • Description.—Chevronlike ribbon trace in convex hyporelief. Specimens 23.5–254 mm long and 2.9–11.2 mm wide. The lobes consist of thin, commalike to arcuate striations or may be plumoselike.

  • Occurrence.—Two lithologies: (1) laminated light to dark gray silty shale; and (2) dark gray (weathered to tan) to pale greenish gray, calcareous silty to sandy shale.

  • Associated ichnotaxa.—Aulichnites isp., Dimorphichnus isp., Diplichnites cf. binatus, Diplichnites cf. govenderi, Lockeia siliquaria, Phycosiphon incertum, and Treptichnus vagans.

  • Discussion.—Protovirgularia cf. pennatus specimen on IBGS PJ-M-019 has characteristics similar to two ichnospecies of Protovirgularia illustrated by Nara and Ikari (2011, fig. 3): P. dichotoma and P. pennatus (Fig. 17.2–17.4). The Protovirgularia cf. pennatus specimen has a medial ridge (convex hyporelief) along the length of the trace, which becomes more prominent near the open end of the trace, characteristic of most Protovirgularia, including P. dichotoma. The striations that form the lateral lobes are very thin and arcuate and similar to the striations of the P. pennatus that form the lateral appendages. The specimen on KUMIP 204521 is winding and has a plumoselike, arcuate striation pattern similar to specimens of P. pennatus (Uchman, 1998, fig. 67A) and Protovirgularia isp. (Knaust, 2007, fig. 7B).

  • Ichnogenus RUSOPHYCUS Hall, 1852

  • Type ichnospecies.—Rusophycus clavatus Hall, 1852, subsequent designation by Miller (1889, pg. 138).

  • Diagnosis.—Small to large bilobate mounds or depressions with parallel or merged lobes near the posterior; parallel to oblique to transverse striations; however, some specimens may be smooth (Crimes, 1970b, Osgood, 1970; Alpert, 1976a; Fillion & Pickerill, 1990; Keighley & Pickerill, 1996).

  • Discussion.—Seilacher (1970) grouped Rusophycus under Cruziana and argued that they should be considered synonymous due to both being produced by the same tracemaker, trilobites. Seilacher also suggested retaining Isopodichnus—morphologically similar to both Cruziana and Rusophycus—for use as a facies indicator for brackish water. Most workers disagree with the Seilacher (1970) suggestion and maintain Rusophycus and Cruziana as separate ichnogenera (See Cruziana for full discussion p. 13). Similar to Cruziana, Rusophycus ichnospecies are separated primarily by striation pattern, but size, lobe morphology (i.e., orientation, ornamentation, shape), and tracemaker morphologic remnants are other criteria sometimes used (Crimes, 1970b; Osgood, 1970; Seilacher, 1970, 2007).

  • Rusophycus is generally interpreted as a resting or hiding trace (e.g., Crimes 1970b; Osgood, 1970; Seilacher, 1970), but also suggested to be a hunting (e.g., Jensen, 1990; Tarhan, Jensen, & Droser, 2011) or nesting (brooding) trace (e.g., Fenton & Fenton, 1937d). Tracemakers of Rusophycus are commonly interpreted as arthropods, such as trilobites and crustaceans, but gastropods, and even some vertebrates have been proposed (e.g., Crimes, 1970a; Seilacher, 1970; Bromley & Asgaard, 1979; Seilacher, 2007). Post-Triassic Rusophycus are not considered produced by trilobites (Fillion & Pickerill, 1990). Jones (2016) showed small bilobate modern bat manus and pes track impressions similar to small, smooth Rusophycus (e.g., R. carbonarius), meaning that bats may have produced some Rusophycus in Cenozoic water-margin environments (e.g., fluvial, lake plain, crevasse-splay deposits). Rusophycus is a facies-crossing ichnogenus reported from shallow marine (e.g., intertidal, lagoon), deep marine (e.g., slope, basin), brackish, lacustrine, and fluvial deposits (e.g., Crimes, 1970b; Seilacher, 1970; Hakes, 1976, 1985; Bromley & Asgaard, 1979; Pollard, 1985; Pickerill, 1995; Garvey & Hasiotis, 2008; Jackson, Hasiotis, & Flaig, 2016). Rusophycus ranges from the Cambrian to recent (e.g., Crimes, 1987; Hasiotis, 2012).

  • Figure 17.

    Protovirgularia specimens from the Spence Shale. 1, Protovirgularia dichotoma in convex hyporelief, KUMIP 314233; 2–4, Protovirgularia cf. pennatus in concave epirelief (2–3), in convex hyporelief (4) with Diplichnites cf. govenderi, KUMIP 204521 A and B; scale bar in cm.

    f17_01.jpg

    RUSOPHYCUS CARBONARIUS (Dawson, 1864)
    Emended by Keighley & Pickerill, 1996
    Figure 18.1–18.3, 18.5

  • Material.—KUMIP 204523 A and B (part and counterpart); 39 specimens, Miner's Hollow; KUMIP 314222 B: one specimen, Miner's Hollow; KUMIP 314223: two specimens, Antimony Canyon; KUMIP 314228: eight specimens, Miner's Hollow; KUMIP 314229: three specimens, Miner's Hollow; IBGS LG-M-011: one specimen; IBGS PJ-M-007: 13 specimens, Miner's Hollow; IBGS PJ-M-008: one specimen, Miner's Hollow; IBGS PJ-M-013: one specimen, Miner's Hollow; IBGS PJ-M-018: two specimens, Miner's Hollow; IBGS PJ-M-023: two specimens, Miner's Hollow.

  • Diagnosis.—Small, bilobate depressions (concave epirelief) or mounds (convex hyporelief) with parallel to slightly gaping lobes; transverse to oblique, fine striations that do not extend beyond the lobe margin, or may be smooth (modified from Keighley & Pickerill, 1997, 1998).

  • Description.—Small bilobate depressions and mounds with a central furrow, typically smooth but may have fine, oblique striations. Burrows 3.0–11.2 mm long and 2.5–6.5 mm wide at the widest point. Only one specimen has fine striations with a 97–120° V-shaped angle.

  • Occurrence.—Two lithologies: (1) gray (weathered to brown), laminated calcareous sometimes with brown siliciclastic sand; and (2) gray, siliciclastic silty shale, sometimes with brown carbonate sand.

  • Associated ichnotaxa.—Bergaueria hemispherica, Cruziana barbata, C. problematica, Gyrophyllites kwassizensis, Lockeia siliquaria, Monomorphichnus lineatus, M. cf. multilineatus, Planolites annularis, P. beverleyensis, P. montanus, Rusophycus cf. cerecedensis, Sagittichnus lincki, Treptichnus bifurcus, and T. vagans.

  • Discussion.—For a full list of synonymy, refer to Keighley and Pickerill (1996). Most Rusophycus specimens from the Spence Shale were assigned to R. carbonarius due to their small size, coffee bean-like shape, and the smoothness of the paired lobes. Normally, the main criterion for classifying ichnospecies of Rusophycus is the surficial striations present on the lobes. Rusophycus carbonarius is characterized by thin, transverse to oblique striations, but Keighley and Pickerill (1996) included small, smooth forms under R. carbonarius because they noted smooth and striated forms on the same samples. Keighley and Pickerill (1997) suggested that the difference between the two forms was taphonomic, rather than ethologic. Supporters of the Seilacher (1970) suggestion to include Rusophycus under Cruziana and the retention of Isopodichnus would likely identify R. carbonarius as Isopodichnus problematicus due to its small size and the lack of striations.

  • Rusophycus carbonarius specimens present in the Spence Shale are significantly smaller than any C. problematica specimen. This is most noticeable on KUMIP 204523 and PJ-M-007 (see Fig. 10.1), where significant differences in width between the two ichnogenera suggest that the same organism did not produce the two ichnofossils (sensu Fortey & Seilacher, 1997). We propose that agnostoid trilobites or small (juvenile) polymeroid trilobites likely produced the Rusophycus carbonarius specimens, whereas medium-sized (adult) polymeroid trilobites likely produced C. problematica specimens. Specimens are oriented in a nearly single direction between 315–350° (relative to the longer cut side of KUMIP 204523B), whereas the C. problematica show an overlapping, curvilinear pattern, suggesting that the bottom currents were relatively strong for smaller organisms and the R. carbonariustracemakers had to orient themselves to the current to remain stable (sensu Pickerill, 1995). Pickerill (1995) illustrated multiple oriented Rusophycus and interpreted their alignment was due to maintaining a rheotactic orientation in waters with significant bottom currents. Multiple specimens of R. carbonarius crosscut several C. problematica specimens suggesting that the R. carbonarius tracemakers may have occupied the area of KUMIP 204523 after the excavation of the C. problematica.

  • RUSOPHYCUS cf. PUDICUS Hall, 1852
    Figure 18.4

  • Material.—IBGS PJ-M-009: one specimen, Spence Shale float, High Creek Canyon, Bear River Range, Utah, USA

  • Diagnosis.—Small- to medium-sized bilobate depressions (concave epirelief) or mounds (convex hyporelief) with parallel to slightly gaping lobes, which widen anteriorly, and transverse to oblique, fine to well-developed striations; the medial furrow well developed and increases in depth and width toward one end, generally extending to entire length of the trace (Osgood, 1970; Fillion & Pickerill, 1990).

  • Description.—Convex, bilobate hyporelief mound with a central furrow, an anterior gape, and a merged posterior. Specimen 14.6 mm long, 9.7 mm wide, and 4.3 mm deep. Central furrow 1.4 mm wide.

  • Occurrence.—Tan to brown, siliciclastic shale with brown, dark gray, or black dendrites.

  • Associated ichnotaxa.—Sagittichnus lincki and Treptichnus bifurcus.

  • Discussion.—Only a single specimen of Rusophycus cf. pudicus was found in the Spence Shale. Like specimens of R. carbonarius, the R. cf. pudicus specimen has smooth lobes. The assignment to R. cf. pudicus is based on the depth increase of the medial furrow, length of furrow equaling the length of the trace, and the wide, well-developed lobes that taper to one end. Within the anterior gape of the central furrow, there is a raised area that may be a poorly preserved impression of the tracemaker coxa. The R. pudicus specimens illustrated by Osgood (1970) were much larger than the specimens shown here, but the size difference in the Spence Shale material could be due to decreased oxygenation or just a smaller tracemaker.

  • RUSOPHYCUS cf. CERECEDENSIS Crimes & others, 1977
    Figure 15.5, Figure 18.6

  • Material.—KUMIP 314228: two specimens, Miner's Hollow; IBGS PJ-M-023: one specimen, High Creek Canyon, Bear River Range, Utah, USA.

  • Diagnosis.—Medium-sized bilobate mound (convex hyporelief); lobes may be rounded or tapered to points and gape anteriorly; individual lobes may be smooth or with oblique to transverse striations.

  • Description.—Convex bilobate hyporelief mounds 12.5–28.9 mm long, 10.4–15.4 mm wide, and 1.5 mm deep. Medial furrow 1.9 mm wide but widens to 6.5 mm anteriorly with a ∼60° V-shaped angle. Oblique striations form ∼100° V-shaped angles.

  • Occurrence.—Brown to gray siliciclastic silty to sandy shale.

  • Associated ichnotaxa.—Cruziana problematica, Lockeia siliquaria, Monomorphichnus lineatus, M. cf. multilineatus, Planolites beverleyensis, P. montanus, and R. carbonarius.

  • Discussion.—Specimens assigned herein to Rusophycus cf. cerecedensis are distinguished from R. carbonarius and R. cf. pudicus by their larger size and rounded to tapered lobe shape. One specimen possibly records a predator-prey interaction with P. beverleyensis (see Fig. 15.5), similar to the association of R. carbonarius and P. montanus (see Fig. 18.3) and to other such associations of Rusophycus and simple burrows (e.g., Helminthopsis, Palaeophycus, or Planolites) (e.g., Jensen 1990; Tarhan, Jensen, & Droser, 2011). Another specimen has a P. montanus burrow oriented at an oblique angle on one lobe (Fig. 18.6), but since the burrow extends into the surrounding host rock, the association is likely not predation.

  • Figure 18.

    Rusophycus specimens from the Spence Shale. 1, Overlapped, individual R. carbonarius (arrows) forming pseudoribbonlike morphology, KUMIP 204523 A and B, Miner's Hollow; 2, Rusophycus carbonarius in convex hyporelief, IBGS PJ-M-007, Miner's Hollow; 3, Planolites montanus terminating at a R. carbonarius in convex hyporelief, IBGS PJ-M-023, Miner's Hollow; 4, Rusophycus cf. pudicus in convex hyporelief, IBGS PJ-M-009, Miner's Hollow; 5, Rusophycus carbonarius with faint striations (circle) on one lobe in concave epirelief, IBGS PJ-M-023, Miner's Hollow; 6, Rusophycus cf. cerecedensis with P. montanus on lobe (arrow), IBGS PJ-M-023, Miner's Hollow; scale 1–4, 6 in cm; 5 in mm.

    f18_01.jpg

    Ichnogenus SAGITTICHNUS Seilacher, 1953b

  • Type ichnospecies.—Sagittichnus lincki Seilacher, 1953b.

  • Diagnosis.—Small, subcircular to ovoid to arrowhead-shaped, convex mounds (hyporelief) or concave pits (epirelief), usually with medial keel; occurring in small to large groups; medial keel may or may not be present (Häntzschel, 1975; Głuszek, 1995; Garvey & Hasiotis, 2008).

  • Discussion.—Sagittichnus is described as small, keeled arrowheadshaped pits and mounds that are usually interpreted as resting traces of an unknown tracemaker (Seilacher, 1953b; Głuszek, 1995; Garvey & Hasiotis, 2008). Bromley and Asgaard (1979) reported specimens from Triassic fresh to brackish lacustrine deposits of Greenland that resembled Sagittichnus but interpreted them as inorganic tool marks, and thus invalid; however, other authors disagree and maintain Sagittichnus as a valid ichnogenus (e.g., Głuszek, 1995; Garvey & Hasiotis, 2008). A recent neoichnological study by Retrum, Hasiotis, and Kaesler (2011) showed freshwater ostracodes producing Sagittichnus-like morphologies. Sagittichnus has also been associated with small arthropod trackways (Głuszek, 1995). Sagittichnus is similar to manus and pes track impressions of modern bat trackways (Jones, 2016). Sagittichnus may occur with or grade into deposit feeding, hiding, or resting traces like Rusophycus (Garvey & Hasiotis, 2008). Sagittichnus has been reported from shallow marine and freshwater to brackish continental deposits (e.g., estuarine, fluvial, lacustrine) (e.g., Bromley & Asgaard, 1979; Głuszek, 1995; Garvey & Hasiotis, 2008; Jackson, Hasiotis, & Flaig, 2016). Sagittichnus ranges from the Cambrian to recent (Bednarczyk & Przybyołwicz, 1980; Retrum, Hasiotis, & Kaesler, 2011; Jackson, Hasiotis, & Flaig, 2016).

  • SAGITTICHNUS LINCKI Seilacher, 1953b
    Figure 13.4, Figure 19.1–19.2

  • Material.—IBGS LG-M-002: five specimens, Winter Hollow, Box Elder Mountain; IBGS LG-M-003: 20 specimens; IBGS LG-M-013: 15 specimens, Antimony Canyon; IBGS PJ-M-025: 15 specimens, Spence Shale Float, Cataract Canyon.

  • Diagnosis.—Small, subcircular to ovoid to arrowhead-shaped, convex mounds lacking discrete medial keels (Garvey & Hasiotis, 2008).

  • Description.—Small, convex hyporelief mounds without medial keel or furrow. Specimens 1.0–2.7 mm long and 0.7–4.6 mm wide.

  • Occurrence.—Tan to dark gray, carbonate or siliciclastic shale.

  • Associated ichnotaxa.—Bergaueria hemispherica, Gyrophyllites kwassizensis, Lockeia siliquaria, Teichichnus cf. nodosus, and Treptichnus bifurcus.

  • Discussion.—Specimens were assigned to Sagittichnus due to their small, subrounded to ovoid-shaped, convex-mound (hyporelief) morphology, and their highly concentrated groupings (Fig. 19.1–19.2). No specimen had the characteristic medial keel (finlike structure) preserved in either epi- or hyporelief. Most specimens are ovoid in shape, but some show a subrounded to arrowhead shape. Also present on IBGS LG-M-003, alongside some S. lincki, is an ovoid-shaped, convex mound that we consider Lockeia for its noticeably larger size than the surrounding Sagittichnus, the lack of a medial keel, and its tapered ends.

  • Ichnogenus SCOLICIA de Quatrefages 1849

  • Type ichnospecies.—Scolicia prisca de Quatrefages, 1849.

  • Diagnosis.—Variable and selectively preserved, simple, winding to meandering to coiling, bilobate or trilobate backfilled burrows; may have one or two parallel, locally discontinuous strings along base; area between strings flat to slightly convex; cross sections circular to oval; geopetal meniscate backfill common but massive burrow infill also common (Häntzschel, 1975; Uchman, 1995).

  • Discussion.—There are many ichnogenera with morphologies similar to Scolicia, informally grouped in the Scolicia Group by Häntzschel (1975, p. 106). Many ichnotaxa from the Scolicia Group were later synonymized with Scolicia (e.g., Uchman, 1995). Plaziat and Mahmoudi (1988) suggested restricting Scolicia to concave epirelief expressions and retaining Subphyllochorda Götzinger & Becker, 1932 for convex hyporeliefs of echinoid traces; however, this complicates ichnotaxonomy more than it helps, and thus, subsequent authors have rejected this suggestion (e.g., Uchman, 1995, 1998; Fu & Werner, 2000).

  • Scolicia is commonly interpreted as a locomotion or depositfeeding trace (e.g., Fu & Werner, 2000); however, some authors have interpreted Scolicia to be a grazing trace (e.g., Uchman, 1995). Scolicia is commonly interpreted as the product of irregular echinoids in the Mesozoic and Cenozoic (Plaziat & Mahmoudi, 1988; Uchman, 1995, 1998), whereas Paleozoic producers were likely gastropods (e.g., Götzinger & Becker, 1932; Häntzschel, 1975; Książkiewicz, 1977). In continental environments since the Devonian, producers were also likely gastropods (e.g., Hasiotis, 2004, 2008; Ash & Hasiotis, 2013). Scolicia has been reported from shallow marine as well as deep marine deposits, including turbidites (Uchman, 1995; Fu & Werner, 2000); however, Fu and Werner (2000) suggested that most shallow marine Scolicia are commonly destroyed by overprinting of deep-penetrating traces. Scolicia tracemakers preferred fine sandy to coarse silty settings, suggesting a preference for lower energy environments (Fu & Werner, 2000). Scolicia ranges from the Cambrian to recent (e.g., Häntzschel, 1975; Fu & Werner, 2000).

  • SCOLICIA isp.
    Figure 20.1–20.6

  • Material.—IBGS PJ-M-032: four specimens, Miner's Hollow, Wellsville Mountains, Utah, USA.

  • Diagnosis.—Short to elongated, cylindrical to subcylindrical burrows in endorelief; undertrace in concave epirelief and convex hyporelief, some may be bilobate with basal medial furrow.

  • Description.—Light to medium brown to gray burrows 17.6– 33.1 mm wide, 5.0–14.6 mm thick, with dark gray burrow margins 0.8–2.7 mm thick. Burrow infills are subangular to subrounded, moderately well-sorted, fine to medium carbonate sand with small reddish brown to red grains and large, very euhedral, dark grains with penetration twinning.

  • Occurrence.—Light to dark gray, laminated siliciclastic mudstone. Laminations are < 3 mm thick. Soft-sediment deformation is present locally around the burrow with flame structures penetrating or deforming the burrow margin. Laminations above and below the Scolicia isp. lack significant bioturbation but several small burrows are present indicating an ii2, whereas the layer with the Scolicia specimen has an ii4–5.

  • Associated ichnotaxa.—None.

  • Discussion.—Specimens assigned to Scolicia (Fig. 20.1–20.2) lack the diagnostic basal bilobate or trilobate shape or double drainage furrows. Cross sections reveal four elliptical burrows with irregularly shaped margins filled with a light to medium brown to gray, fine to medium sand in a matrix of laminated, light gray silt to fine sand (Fig. 20.3–20.6). The burrow margins are composed of dark gray, fine to medium sand (Fig. 20.3). One burrow appears bilobate from presence of a possible medial furrow composed of a wedge of light gray mud partially separating the burrow into two lobes (Fig. 20.4); however, the medial ridge may be due to compaction and soft-sediment deformation (e.g., flame structures) as other burrows have similar structures penetrating them from the sides. Also present is possible fecal-drainage canal near the base of one lobe, formed by a circle of dark sand grains with a brown core (Fig. 20.4). The irregularity of the dark burrow margins may also be the result of soft-sediment deformation and postdepositional diagenesis. The burrow infill has multiple coarse, angular, dark grains that may have resulted from recrystallization during diagenesis, as some of the grains are very euhedral and one grain appears to exhibit penetration twinning (Fig. 20.5–20.6).

  • Figure 19.

    Sagittichnus, Taenidium, and Teichichnus specimens from the Spence Shale. 1, Small field of Sagittichnus lincki, convex hyporelief, IBGS PJ-M-025; 2, field of Sagittichnus lincki (black arrows) with Lockeia siliquaria (white arrow) in convex hyporelief, IBGS LG-M-003; 3, Taenidium cf. satanassi crosscut by Archaeonassa jamisoni isp. nov. (black arrows) and insertion furrow (white arrow) in partial endorelief, IBGS PJ-M-002, Miner's Hollow; 4, Taenidium cf. satanassi with meniscate backfill (arrows) in partial endorelief, IBGS PJ-M-002, Miner's Hollow; 5, Segmented Teichichnus cf. nodosus in convex hyporelief, IBGS PJ-M-029; 6, Cross section of Te. cf. nodosus (5) showing characteristic gutter-shaped spreite; scale 1–5 in cm; 6 in mm.

    f19_01.jpg

    Ichnogenus TAENIDIUM Heer, 1877

  • Type ichnospecies.—Taenidium serpentinum Heer, 1877.

  • Diagnosis.—Unlined to thinly lined, unbranched, straight to sinuous, cylindrical burrows with meniscate segmented burrow fill (D'Alessandro & Bromley, 1987).

  • Discussion.—Prior to D'Alessandro and Bromley (1987) reexamining the original descriptions and type material of Muensteria Sternberg, 1833 and Taenidium Herr, 1877, most workers used Muensteria for unbranched, unlined meniscate burrows, whereas Taenidium was used for branching meniscate burrows. Muensteria was considered invalid as the original description was confusing and included algae, coprolites, and several forms of Chondrites Sternberg, 1833 (D'Alessandro & Bromley, 1987). Taenidium was recommended for unbranched meniscate burrows previously described as Muensteria and a new ichnogenus, Cladichnus, was erected for meniscate burrows with primary successive branching or radiating systems (D'Alessandro & Bromley, 1987).

  • Keighley and Pickerill (1994) reviewed Beaconites Vialov, 1962, and compared it to other meniscate-backfilled burrows, Ancorichnus Heinberg, 1974, and Taenidium. They considered Beaconites barretti Bradshaw, 1981, as an unlined, unwalled meniscate burrow belonging to Taenidium and argued that the ends of the menisci do not form a wall or lining. Many authors followed Keighley and Pickerill (1994) for the use of Taenidium barretti (e.g., Schlirf, Uchman, & Kümmel, 2001; Keighley & Pickerill, 2003; Buatois & others, 2007).

  • Beaconites barretti is valid and still retained by many authors (e.g., Morrissey & Braddy, 2004; Smith & Hasiotis, 2008; Smith & others, 2008b; Counts & Hasiotis, 2009) because its architectural morphology is clearly distinct from Taenidium, rejecting the synonymy of most backfilled burrows into Taenidium by Keighley and Pickerill (1994). Beaconites is an unlined, tightly spaced backfilled meniscate burrow where the backfills merge laterally to form a crenulated burrow wall, representing the remnant of an open cell as it was moved through the sediment. We find the Keighley and Pickerill (1994) definitions of walls and linings to be confusing and inappropriate to all backfilled-burrow morphologies. Keighley and Pickerill (1994) considered backfilled burrows to not have true walls or linings as they considered simple excavation to not be a form of active construction, their requirement for walls and linings. They also interchanged the terms wall and lining, causing their definitions and usage to become muddled. Linings are only one possible type of wall structure (sensu Bromley, 1996), whereas Keighley and Pickerill (1994, fig. 1) considered all walled ichnofossils to have linings or mantles. A wall is the outermost margin of the area the tracemaker occupied—regardless of its active or passive excavation or construction (contra Keighley & Pickerill, 1994)—where the burrow infill contacts the matrix (sensu Morrissey & Braddy, 2004). Smith and others (2008b) argued that the overlapping of menisci form a crenulated, but unlined, wall in B. barretti reflecting active excavation of the sediment by the tracemaker, relocating it to the rear of the active cell, and compacting it to form the rear wall. We, therefore, follow Smith and others (2008b) for the retention of Beaconites barretti and definitions of walls vs. lining. Taenidium should be restricted to burrows that exhibit thick, regularly spaced meniscate backfill that is symmetrical about the axis of the burrow, which is unlined and unbranched (D'Alessandro & Bromley, 1987; Smith & others, 2008b). Prior to the inclusion of B. barretti by Keighley and Pickerill (1994), Taenidium was only described from marine deposits. Taenidium reported from continental deposits (e.g., Savrda & others, 2000; Buatois & Mángano, 2002, 2007, 2011; Krapovickas & others, 2009; Scott & Smith, 2015) actually belong to: (1) Naktodemasis Smith & others, 2008b, if the thin meniscate backfill are organized into discreet packages; (2) Beaconites, if the menisci are uneven, alternate around a central axis, and not organized into discreet packages; or (3) Ancorichnus Heinberg, 1974, if a mantle is present (e.g., Smith & others, 2008b; Counts & Hasiotis, 2009; Morshedian, MacEachern, & Dashtgard, 2012; Gingras & others, 2016; Harris & others, 2016).

  • Taenidium is interpreted as a deposit-feeding trace of marine worms (Gevers & others, 1971; Keighley & Pickerill, 1994; Smith & others, 2008b). Taenidium has been reported from shallow- to deep-marine deposits (Keighley & Pickerill, 1994, fig. 5; Smith & others, 2008b; Jackson, Hasiotis, & Flaig, 2016). Taenidium has been reported from the Vendian (i.e., Ediacaran) by Germs (1972) and Jenkins (1995); however, Jensen, Droser, and Gehling (2006), considered them as a cast of Cloudina and a tubular fossil, respectively. We follow the interpretation of Germs (1972) and Jenkins (1995) based on the similarity of the morphologies to Taenidium. Therefore, Taenidium ranges from the Ediacaran to recent (e.g., Germs, 1972; Crimes, 1992; Jenkins, 1995; Uchman, 1998; Jackson, Hasiotis, & Flaig, 2016).

  • TAENIDIUM cf. SATANASSI D'Alessandro & Bromley, 1987
    Figure 19.3–19.4

  • Material.—IBGS PJ-M-002: one specimen, Miner's Hollow; IBGS PJ-M-005: one specimen, Spence Shale float, Miner's Hollow.

  • Diagnosis.—Long, slightly sinuous to straight burrow with uniform, evenly spaced, meniscate backfill; meniscate packages shorter than burrow diameter and filled with alternating sediment types (D'Alessandro & Bromley, 1987).

  • Description.—Straight to gently curved, endorelief burrow with gray meniscate backfill and brown to purple weathered infill. Burrows 152.2–159.4 mm long, 6.3 mm wide; burrow menisci ∼1.5 mm thick and uniform.

  • Occurrence.—Two lithologies: (1) tan to light brown, siliciclastic silty shale; and (2) gray, calcareous shale.

  • Associated ichnotaxa.—Archaeonassa jamisoni, Phycodes curvipalmatum, and Planolites; montanus.

  • Discussion.—The long, mostly straight burrow on IBGS PJ-M-002 was assigned to Taenidium cf. satanassi due to the presence of meniscate backfill exposed by a large ovoid depression, herein designated as Archaeonassa jamisoni. Exposed menisci are shorter designated as Archaeonassa jamisoni. Exposed menisci are shorter than the burrow diameter but lack the sediment alternation characteristic of T. satanassi (Fig. 19.3–19.4). The rest of the specimen occurs in endorelief and shows no clear internal structure; however, the purple coloration of the weathered burrow infill has a slight serrated pattern near the burrow margins, possibly a diagenetic remnant of the meniscate backfill. The specimen of Taenidium cf. satanassi on IBGS PJ-M-005 is completely in endorelief, revealing no internal morphology, and is in close proximity to several A. jamisoni specimens.

  • Figure 20.

    Scolicia specimens from the Spence Shale. 1, Plan view of Scolicia isp. specimens with cross-section axis (lines) and weathered edge (arrow), PJ-M-032, Miner's Hollow; 2, weathered edge of specimen with two Scolicia isp. burrows exposed; 3, cross section of burrows near the weathered edge, note irregular burrow margins; 4—6, Scolicia isp. cross-sections with variably colored infill and associated soft-sediment deformation; 4, Scolicia isp. with light tan infill with possible basal medial furrow (arrow) and interpreted fecal drainage canal (box); 5, Scolicia isp. with tan infill, thick dark-gray burrow margin, and coarse euhedral grains (white arrows) with possible twinning (gray arrow); 6, gray-brown Scolicia isp. with thin burrow margin (arrow) and coarse grains.

    f20_01.jpg

    Ichnogenus TEICHICHNUS Seilacher, 1955b

  • Type ichnospecies.—Teichichnus rectus Seilacher, 1955b, by original monotypy.

  • Diagnosis.—Long, wall-shaped, septate structures consisting of stacks of gutter-shaped laminations (Seilacher, 1955b Fillion & Pickerill, 1990).

  • Discussion.—Teichichnus was introduced for vertically stacked, horizontal burrows with spreiten and thought to be produced by upwardly shifting deposit feeders (Seilacher, 1955b Fillion & Pickerill, 1990). Teichichnus has been reported to intergrade with multiple ichnofossils: Cruziana, Ophiomorpha Lundgren, 1891, Phycodes, Rhizocorallium Zenker, 1836, and Thalassinoides Ehrenberg, 1844 (e.g., Fillion & Pickerill, 1990; Loope & Dingus, 1999). Teichichnus has been noted for its similarity to Trichophycus Miller & Dryer, 1878, due to the presence of gutter-shaped spreite (e.g., Osgood, 1970; Frey & Howard, 1985; Geyer & Uchman, 1995), but can be typically distinguished by the more planar shape of the spreite and a lack of fine striations present on the outside of the burrow (e.g., Fillion & Pickerill, 1990; Jensen, 1997). Some Teichichnus have been reported to have surficial striations (e.g., Jensen, 1997).

  • Teichichnus is typically interpreted as a deposit-feeding or grazing trace of annelids and arthropods (e.g., Chisholm, 1970; Fillion & Pickerill, 1990). Teichichnus primarily occurs in shallow-marine deposits (e.g., tidal flats and deltas) but some have been reported from deep-marine (e.g., submarine fans and abyssal plain) and brackish-marine (meso- to polyhaline) water deposits (e.g., Fürsich, 1975; Fillion & Pickerill, 1990; Pemberton & Wightman, 1992; Gingras, MacEachern, & Pemberton, 1998; Jackson, Hasiotis, & Flaig, 2016). Teichichnus ranges from the early Cambrian to recent (e.g., Narbonne & others, 1987; Fillion & Pickerill, 1990; MacNaughton & Narbonne, 1999).

  • TEICHICHNUS cf. NODOSUS Fillion & Pickerill, 1990
    Figure 19.5–19.6

  • Material.—IBGS PJ-M-025: one specimen, Spence Shale Float, Cataract Canyon.

  • Diagnosis.—Large, curved, undulating burrow with spreiten forming chain of irregularly spaced nodes preserved in convex hyporelief (Fillion & Pickerill, 1990).

  • Description.—Curved, undulating, segmented burrow with retrusive spreite. Burrow 73.4 mm long, 7.2–12.5 mm wide, and burrow segments and internodes 2.8–4.3 mm thick. Light to dark gray spreite 0.3–0.6 mm thick, and composed of fine to very fine sand.

  • Occurrence.—Light to dark gray (weathered to tan), laminated fine to very fine sand.

  • Associated ichnotaxa.—Bergaueria hemispherica and Sagittichnus lincki.

  • Discussion.—Teichichnus cf. nodosus (Fig. 19.5) was assigned due to its similarity to the undulating and nodular morphology of T. nodosus Fillion & Pickerill, 1990. This specimen also occurs with a partial eocrinoid, Gogia granulosa Robison, 1965. The specimen of Teichichnus cf. nodosus terminates near a B. hemispherica (see Fig. 6.6). A cross section of T. cf. nodosus reveals several gutter-shaped spreiten that alternate between brown, light gray, and dark gray fine-grained sand (Fig. 19.6).

  • Ichnogenus TREPTICHNUS Miller, 1889
    Emended by Buatois & Mángano, 1993a

  • Type ichnospecies.—Treptichnus bifurcus Miller, 1889 (p. 581).

  • Diagnosis.—Chains of horizontal to subhorizontal, straight to curved, zigzagging burrow segments associated with vertical to oblique tubes producing a three-dimensional burrow structure; pits and nodules may occur near top or base of burrow segments at sediment interfaces (Buatois & Mángano, 1993a; Uchman, Bromley, & Leszczyński, 1998).

  • Discussion.—Miller (1889) named Treptichnus for forked, zigzagging burrows with projected burrow ends; Miller interpreted the burrow projections as indicating the direction of tracemaker movement and were produced by insect larvae or pupa. Along with Treptichnus, Miller (1889) also described and commented on Haplotichnus and Plangtichnus as being very similar to Treptichnus in terms of size, tracemaker, and morphology. Plangtichnus is similar to Treptichnus and was originally described as a zigzag trail with pits deeper than the rest of the trail (Miller 1889, p. 580). Archer and Maples (1984) and Maples and Archer (1987) argued that Plangtichnus is distinguishable from Treptichnus by the lack of burrow-end projections that yields a highly angular zigzagging form; however, Buatois and Mángano (1993a) argued that the projections of Treptichnus and the pits of Plangtichnus represented morphologically similar vertical shafts along the burrow system and that the lack of the burrow-end projections was likely caused by erosion. Buatois and Mángano (1993a) claimed that, since both ichnogenera had similar morphology and represented similar behaviors, Plangtichnus and Treptichnus should be considered synonymous. They retained Treptichnus for nomenclatural stability and considered Plangtichnus to be nomen oblitum, citing relative nonuse of the name. Treptichnus pollardi was, therefore, erected as a new ichnospecies to replace the name for the morphology previously associated with Plangtichnus erraticus (Buatois & Mángano, 1993a).

  • Treptichnus is commonly interpreted as a deposit-feeding trace (Buatois & Mángano, 1993a, 1993b; Uchman, Bromley, & Leszczyński, 1998), but has also been interpreted to be an agricultural, grazing, reproduction, and predation or scavenging trace (e.g., Rindsberg & Kopaska-Merkel, 2005; Seilacher, 2007; Vannier & others, 2010; Wilson & others, 2012; Getty & others, 2016). Treptichnus is interpreted as being produced by marine annelid worms (e.g., Buatois & Mángano, 1993a, 1993b; Uchman, Bromley, & Leszczyński, 1998; Vannier & others, 2010) and some insect larvae in continental environments since the Pennsylvanian (e.g., Miller, 1889; Rindsberg & Kopaska-Merkel, 2005; Getty & others, 2016). Treptichnus has been reported from shallow- and deep-marine, and continental proximal floodplain and proximal lacustrine deposits (e.g., Archer & Maples, 1984; Buatois & Mángano, 1993b; Jensen, 1997; Uchman, Bromley, & Leszczyński, 1998; Wilson & others, 2012; Getty & others, 2016). Treptichnus ranges from the Cambrian to recent (e.g., Buatois & Mángano, 1993a, 1993b; Uchman, Bromley, & Leszczyński, 1998; Vannier & others, 2010; Hasiotis, 2012); however, some Treptichnus have been reported from the Edicaran and were suggested to represent a gradual increase in ichnofossil complexity until the first occurrence of T. pedum at the Precambrian-Cambrian boundary (e.g., Germs, 1972; Jensen & others, 2000; Gehling & others, 2001; Droser & others, 2002).

  • TREPTICHNUS BIFURCUS (Miller, 1889)
    Figures 10.5; 13.5; 21.1–21.6; and 22.1–22.3

  • Material.—KUMIP 204523 A+B; one specimen, Miner's Hollow; KUMIP 314230; one specimen, Antimony Canyon; KUMIP 314250; three specimens, Miner's Hollow; KUMIP 314283: one specimen; IBGS PJ-M-006; one specimen, Miner's Hollow; IBGS PJ-M-008: one specimen, Miner's Hollow; IBGS PJ-M-009: one specimen, Spence Shale float, High Creek Canyon, Bear River Range, Utah, USA; IBGS PJ-M-028: one specimen, Spence Shale Float, Miner's Hollow; IBGS PJ-M-030 (part and counterpart): one specimen, Miner's Hollow.

  • Diagnosis.—Burrow system with short projections between elongate, thin, and horizontal burrow segments forming straight to slightly curved, zigzagged chains; may occur as chains of evenly spaced beads or depressions alternating around central axis, forming zigzag pattern (Buatois & Mángano, 1993a; Uchman, Bromley, & Leszczyński, 1998).

  • Description.—A zigzag-segmented burrow system 30.2–120.3 mm long, 8.6–18.4 mm wide with burrow projections. Segments 7.4–33.2 mm long, 1.1–4.7 mm wide; circular to subrounded, depression or bead diameter 1.9–3.9 mm, nonalternating beads spaced 10.7–23.7 mm; whereas alternating beads spaced 5.5–21.0 mm. Angles between burrow segments range from 66–129°, average 99°. Specimens occur in concave and convex hyporelief and epirelief.

  • Occurrence.—Thickly laminated to massive, medium to dark gray or tan to light brown, siliciclastic silty or calcareous shale.

  • Associated ichnotaxa.—Cruziana problematica, Gyrophyllites kwassizensis, Lockeia siliquaria, Monomorphichnus lineatus, M. cf. multilineatus, Planolites beverleyensis, P. montanus, Rusophycus carbonarius, Rusophycus cf. pudicus, Sagittichnus lincki, and Treptichnus vagans.

  • Discussion.—Treptichnus bifurcus is the most common form of Treptichnus from the Spence Shale. Specimens exhibit two primary morphologies with most occurring as chains of simple, short, straight zigzagging burrow segments with short projections of the older segment past the start of new segment (Fig. 21.1, 21.3–21.5). The other T. bifurcus morphology has more curved or slightly meandering burrow segments (Fig. 21.2, 21.6). The projections at the end of burrow segments have been interpreted as compressed portions of the oblique shafts (Maples & Archer, 1987). Getty and others (2016), however, recently argued that burrow projections in Treptichnus were not formed by compression and resulted from the tracemaker backing into the previous segment, changing directions, and constructing a new segment within the same plane. Treptichnus bifurcus is one of the few traces previously reported from the Spence Shale by Robison (1969, pl. 138, fig. 5) as “burrow type A” and “feather-stitch burrow”. The term “feather-stitch trail” was widely used in the literature prior to the 1970s before the rediscovery of the Miller (1889) paper (Buatois & Mángano, 1993a; Uchman, Bromley, & Leszczyński, 1998).

  • Alternating beaded T. bifurcus specimens (Fig. 22.1–22.3) are similar to the upper surface features of T. bifurcus and T. pollardi in Buatois and Mángano (1993a, fig. 2B, 3B). Reconstructions of T. bifurcus and T. pollardi show both ichnospecies may occur as a series of pits alternating along a central axis in the upper portions of Treptichnus systems and were interpreted as the burrow apertures of vertical to oblique shafts (Buatois & Mángano, 1993a). Since both T. bifurcus and T. pollardi may occur as alternating pits, assignment of alternating beaded specimens to any one Treptichnus ichnospecies is usually not possible. Specimens present on IBGS PJ-M-006, however, occur in very close proximity to a long T. bifurcus specimen with similar diameters of burrow segments, suggesting the specimen could be part of the T. bifurcus and, thus, included within the type ichnospecies.

  • Alternating beaded T. bifurcus specimens also bear a resemblance to Treptichnus isp. 5 from Buatois and Mángano (1993a, fig. 4), which also occurs as a chain of alternating pits. Treptichnus isp. 5 pits, however, are connected into pairs by burrow segments that do not connect to another pit-burrow segment pair, whereas alternating beaded T. bifurcus specimens are not connected into pairs. Treptichnus specimens from the Eocene Green River Formation (Hogue & Hasiotis, in review) share the alternating beaded T. bifurcus morphology and grade into a single-chain, beaded morphology, which in turn grades into a pitted furrow Ptychoplasma (Protovirgularia) vagans-like morphology (for discussion see Treptichnus vagans p. 43).

  • TREPTICHNUS PEDUM (Seilacher, 1955b)
    Figure 23.1

  • Material.—IBGS PJ-M-017: one specimen, Spence Shale Float, Miner's Hollow; IBGS PJ-M-027: one specimen, Miner's Hollow.

  • Diagnosis.—Treptichnus burrow system consisting of subhorizontal, straight to curved primary burrow with multiple successive burrow segments branching off in regular intervals (Fillion & Pickerill, 1990; Jensen, 1997).

  • Description.—Winding burrow system with systematic projection of burrow segments from a primary burrow. Burrow system 26.6–131.5 mm long, 10.2–18.8 mm wide. Burrow segments 6.4–41.8 mm long, 0.6–4.9 mm wide.

  • Occurrence.—Laminated light gray and medium gray or medium gray and dark gray calcareous silty shale.

  • Discussion.—Originally, the epithet “pedum” was assigned to Phycodes by Seilacher (1955b) for a system of burrow segments that successively branch off along a primary tunnel. Jensen (1997) transferred Phycodes pedum to Treptichnus (see Jensen, 1997 for discussion). Geyer and Uchman (1995) transferred P. pedum to Trichophycus due to the presence of Teichichnus-like spreiten in some burrow segments; however, most authors currently follow Jensen (1997) on the use of Treptichnus pedum (e.g., Jensen & others, 2000; Seilacher, 2007; Wilson & others, 2012; Buatois, Almond, & Germs, 2013). In an attempt to make ichnotaxonomy follow the rules of parsimony common in other areas of science, Dzik (2005) proposed that ichnofossils should be viewed as body fossils and split Treptichnus and placed T. pedum into one of two new worm genera, Manykodes. We disagree with the Dzik (2005) proposal, as parsimony is not always applicable to ichnotaxonomy and to consider ichnofossils as biological taxa would greatly diminish their usefulness in sedimentology and stratigraphy.

  • Treptichnus pedum specimens occur as convex hyporeliefs on samples IBGS PJ-M-017 and IBGS PJ-M-027. On IBGS PJ-M-017, T. pedum occurs in hyporelief and most burrow segments are convex, whereas others are concave (Fig. 23.1). Treptichnus pedum also occurs with some specimens of Cruziana problematica. The burrow segments are elongated and straight to curved extending from a master tunnel (Maples & Archer, 1987). Some of the straighter segments widen at one end, which suggest the segments were oriented obliquely to bedding and later flattened during compaction like in specimens of T. bifurcus.

  • Figure 21.

    Treptichnus bifurcus specimens from the Spence Shale. 1, Convex hyporelief, KUMIP 314230, Antimony Canyon; 2–3, convex hyporelief (2) and concave hyporelief (3), KUMIP 314230, Miner's Hollow; 4, concave epirelief, KUMIP 314283; 5, concave epirelief, IBGS PJ-M-028, Miner's Hollow Float; 6, convex epirelief with yellow-brown burrow infill (black arrow) and concave epirelief Rusophycus carbonarius (white arrow), IBGS PJ-M-008, Miner's Hollow; scale bars in cm.

    f21_01.jpg

    Figure 22.

    Treptichnus bifurcus specimens from the Spence Shale (continued). 1, Treptichtius bifurcus with eroded shaft bases (circles), IBGS PJ-M-006, Miner's Hollow; 2–3, Eroded shaft bases of T. bifurcus (circles) with trace axis (line), specimens of coprolite chain (black arrows), and Planolites montanus (white arrow) in convex hyporelief, IBGS PJ-M-030, Miner's Hollow; 4–5, Jellyfish? impression with Elrathia? sp. trilobite in part (4) in convex epirelief and counterpart (5) in concave hyporelief, KUMIP 314121; scale bars in cm.

    f22_01.jpg

    TREPTICHNUS VAGANS (Książkiewicz, 1977)
    Figure 23.2–23.5, Figure 24.1–24.6

  • Thin, threadlike discontinuous trails–Germs, 1972, p. 866, pl. 1, fig. 5, 7, pl. 2, fig. 1.

  • *Tuberculichnus vagansKsiążkiewicz, 1977, p. 140, pl. 13, fig. 4, text-fig. 27C–G.

  • Tuberculichnus meandrinusKsiążkiewicz, 1977, p. 141, pl. 13, fig. 5–6, text-fig. 27A–B.

  • Hormosiroidea canadensisCrimes & Anderson, 1985, p. 325, fig. 8.1, 9.

  • Hormosiroidea arumberaWalter, Elphinstone, & Heys, 1989, p. 244, fig. 14D–E, 15B, 15D.

  • Hostynichnium duplexPlička & Siránová, 1989, p. 110, pl. 63.

  • Tuberculichnus vagans Książkiewicz—Uchman, 1991, p. 209.

  • Tuberculichnus vagans Książkiewicz—Uchman, 1992, p. 432.

  • String pits—Buatois & Mángano, 1993b, p. 246, fig. 4G.

  • non Tuberculichnus vagans Książkiewicz—Löffler & Geyer, 1994, p. 513, fig. 4E (=Margaritichnus or Microspherichnus).

  • Tuberculichnus meandrinus Książkiewicz—Paczesna, 1996, p. 67, pl. 29, fig. 5.

  • Tuberculichnus vagans Książkiewicz—Pacześna, 1996, p. 67, pl. 29, fig. 8, [non pl. 30, fig. 1, 3].

  • Tuberculichnus vagans Książkiewicz—Buatois & others, 1995, p. 268, fig. 6A, B, 7–8.

  • Tuberculichnus vagans Książkiewicz—Buatois & others, 1996, p. 296, fig. 10C–D.

  • Protovirgularia vagans (Książkiewicz)—Uchman, 1998, p. 166, fig. 70.

  • Treptichnus pedum triplex (Seilacher)—Seilacher, 2007, p. 182, pl. 64, fig. A–B.

  • Protovirgularia vagans (Książkiewicz)—Uchman, 2007, p. 230, pl. 3, fig. 10, pl. 4, fig. 1.

  • Tuberculichnus vagans (Książkiewicz)—Uchman, 2008a, p. 64, fig. 120.

  • Protovirgularia vagans (Książkiewicz)—Uchman, 2008b, p. 130, fig. 8.8 B–C.

  • Ptychoplasma vagans (Książkiewicz)—Uchman, Mikuláš, & Rindsberg, 2011, p. 394, fig. 3A–G, 4A–B.

  • Linear rosary structures—Caron and others, 2010, Supplementary material 8, p. 16, fig. DR6 A–B.

  • Rosary-like structures—Mángano, 2011, p. 98, text-fig. 3, 4C–D, 5, 6A–F.

  • Ptychoplasma vagans (Książkiewicz) (as Fenton & Fenton, 1937b)— Alonso-Muruaga, Buatois, & Limarino, 2013, p. 232, fig. 3E.

  • non Ptychoplasma vagans (Książkiewicz)—Paranjape, Kulkarni, & Gurav, 2013, p. 1366, pl. 3, fig. G–I (=Halopoa or Palaeophycus).

  • non Ptychoplasma vagans (Książkiewicz)—Hagdorn, 2014, p. 268, fig. 12.2. (=Lockeia).

  • non Ptychoplasma vagans (Książkiewicz)—Knaust, Warchoł, & Kane, 2014, p. 2252, fig. 6D (=Palaeophycus or Planolites).

  • Ptychoplasma vagans (Książkiewicz)—Stachacz, 2016, p. 316, fig. 17 G.

  • Treptichnus bifurcus Miller—Getty & others, 2016, p. 273, fig. 4.5.

  • Material.—KUMIP 314122: one specimen, Antimony Canyon; KUMIP 314217: one specimen, Miners Hollow; KUMIP 314222 A–C: five specimens, Miner's Hollow; KUMIP 314233: one specimen; KUMIP 314235: one specimen; IBGS LG-M-004: one specimen; IBGS LG-M-012: one specimen; IBGS LG-M-013: two specimens; IBGS PJ-M-008: one specimen, Miners Hollow; IBGS PJ-M-014: one specimen, Miners Hollow; IBGS PJ-M-019: two specimens, Miners Hollow; IBGS PJ-M-031: four specimens.

  • Emended Diagnosis.—Irregularly meandering or looping, discontinuous trail of variably spaced, short to elongate, ovoid to irregular to circular beads (hyporelief) or depressions (epirelief).

  • Description.—Trails 16–314 mm long, may overlap and cross each other. Beads 0.6–7.0 mm long, 0.6–3.3 mm wide, spaced 0.7–7.0 mm apart. Most specimens are convex hyporelief; however, two specimens are concave epirelief (KUMIP 314122 and KUMIP 314235), and one specimens is preserved as part and counterpart (KUMIP 314233).

  • Occurrence.—Three lithologies: (1) gray (weathers to brown), mica-rich silty shale; (2) dark gray calcareous shale; and (3) gray to dark gray siliciclastic shale.

  • Associated ichnotaxa.—Dimorphichnus isp., Diplichnites gouldi, Halopoa aff. imbricata, Lockeia siliquaria, Monomorphichnus bilinearis, Phycosiphon incertum, Planolites montanus, Protovirgularia cf. pennatus, Rusophycus carbonarius, Sagittichnus lincki, and Treptichnus bifurcus.

  • Discussion.—Originally, the epithet “vagans” was assigned to Tuberculichnus by Książkiewicz (1977) for irregularly winding chains of ridgelike knobs. Uchman (1998) moved Tuberculichnus vagans to Protovirgularia for the amygdaloidal shape of said knobs. Uchman, Mikuláš, and Rindsberg (2011) later transferred Protovirgularia vagans to Ptychoplasma Fenton & Fenton, 1937b, for the carinate shape of the knobs. Mángano and others (2002) suggested P. vagans should be considered a form of Lockeia due to its amygdaloidal, carinate shape, and lack any chevronate patterns. Ptychoplasma vagans from Paranjape, Kulkarni, and Gurav (2013) closely resemble Halopoa or Palaeophycus and lack the diagnostic beaded morphology. Hagdorn (2014) illustrated P. vagans specimens occurring in short chains and not winding chains as in the type material. The P. vagans specimens illustrated by Knaust, Warchoł, and Kane (2014) do not form chains, and the ridgelike knobs do not appear connected. The repeated transfer of P. vagans and wide range of reported morphologies has caused a significant problem regarding its identification and ichnotaxonomic status.

  • Herein, we transfer P. vagans to Treptichnus based on morphological similarities to this ichnotaxon in an attempt to stabilize the nomenclature. The morphology of P. vagans is ill suited for inclusion in Protovirgularia, Ptychoplasma, and Lockeia due to the lack of chevronate and bilobate morphology and the beaded morphology that does not match the type material, respectively. Spence Shale specimens are most similar to P. vagans as both share a winding, single-chain beaded morphology in convex hyporelief. Ptychoplasma vagans, however, sometimes forms irregularly shaped furrows in concave epirelief, which some Spence Shale specimens do as well (Fig. 23.2). Similar to both ichnotaxa is a specimen of Treptichnus from the Eocene Green River Formation (Fig. 23.4), which incorporates aspects of Spence Shale chain specimens, alternating beaded Treptichnus bifurcus, and P. vagans to form: (1) a concave epirelief, alternating beaded morphology (=beaded Treptichnus bifurcus); (2) transitions to a beaded singlechain morphology (=beaded Treptichnus pedum); and then to (3) a pitted furrow morphology (=P. vagans sensu Uchman, Mikuláš, & Rindsberg, 2011) (Hogue & Hasiotis, in review). Due to the similar morphology between ichnotaxa, we, therefore, place Ptychoplasma (Protovirgularia) vagans within Treptichnus as a valid ichnospecies, and it should be referred to as Treptichnus vagans. We interpret the nodes or pits at the base or top of vertical to subvertical shafts as similar to those in T. bifurcus and T. pollardi, as well as a Treptichnus where the tracemaker probed through the sediment interface in a relatively straight line. Erosion of the upper part of the trace would leave an apparent string of beads, with or without remnants of the associated shafts.

  • Treptichnus vagans is composed of long, winding trails of beads that vary in shape from circular to ovoid to irregular (Fig. 23.2–23.3,5, Fig. 24.1). Most specimens with circular beads cross themselves (Fig. 23.2,5); whereas, only one specimen with ovoidshaped beads does (Fig. 24.1). The individual bead shape in a single chain is not always morphologically uniform. One specimen with ovoid-shaped beads has short “drag” marks on one end of the beads extending in the same direction, which are interpreted as insertion furrows (see Fig. 23.3). Another specimen has a mix of different bead morphologies: ovoid, fluted (grooves), imbricated, and triangular (Fig. 24.2–24.6). The fluted and triangular bead morphologies may have been produced by protobranch mollusks due to their V shape (sensu Seilacher & Seilacher, 1994).

  • Treptichnus vagans is similar to numerous “string-of-pearls” ichnotaxa, such as Margaritichnus Bandel, 1973, and Microspherichnus Hakes, 1976. Both ichnotaxa consist of long, sometimes meandering, trails of circular to oval-shaped depressions or mounds. Margaritichnus is usually preserved in convex epirelief with mounds closely spaced that are commonly in contact with each other (Hakes, 1976). Microspherichnus is also preserved in convex epirelief with irregularly spaced beads that may or may not be in contact with each other (Hakes, 1976; Fillion & Pickerill, 1990). Treptichnus vagans specimens also resemble the “string pits” of Buatois and Mángano (1993b, fig. 4G). The “string pits” were described as a hypichnial chain of small, subrounded to oval mounds (< 1 mm) spaced 0.5–4.0 mm apart (max. length = 100 mm) and originally interpreted as locomotion traces of an unknown arthropod. The Buatois and Mángano (1993b) “string pits” are included within T. vagans.

  • Some Treptichnus pedum specimens have been reported with a beadlike morphology similar to T. vagans. Treptichnus pedum specimens from the lower Cambrian of Namibia commonly occur in long, sinuous chains with ovoid to subrounded to circular beads (e.g., Germs, 1972; Jensen & others 2000; Seilacher, 2007; Wilson & others, 2012). Germs (1972) described long, sinuous chains of discontinuous ridges that were later regarded as Treptichnus pedum triplex by Seilacher (2007). The discontinuous ridges of Germs (1972) are almost identical to the Książkiewicz (1977) type material and are included in T. vagans. Multiple specimens of T. pedum with subrounded to circular beads were reported from the lower Cambrian of Namibia (Wilson & others, 2012, fig. 10–12). Some of the Namibian T. pedum specimens with beads are amalgamated together to form recognizable burrow segments (Wilson & others, 2012, fig. 12e–g). Jensen and others (2000) also presented chains of beaded trails assigned to T. pedum that may be better assigned to T. vagans. The Jensen and others (2000) and Wilson and others (2012) specimens likely represent an intergradation between T. pedum and T. vagans.

  • Treptichnus vagans specimens are similar to Hormosiroidea canadensis Crimes & Anderson, 1985 and H. arumbera Walter, Elphinstone, & Heys, 1989. Uchman (1995) later transferred H. canadensis to Saerichnites Billings, 1866, arguing that the vertical-tube expression was inconsistent with the diagnosis and type ichnospecies of Hormosiroidea. Hormosiroidea Schaffer, 1928, is characterized as a horizontal chain of spheres or depressions connected by a central tube, whereas Saerichnites was established as a trackway of paired, parallel rows of alternating semicircular to subquadrate pits (Häntzschel, 1975). Crimes and Anderson (1985) considered the beads and depressions of H. canadensis to be expressions of a vertical meandering method or vertical shafts that were connected by a horizontal tube. Walter, Elphinstone, and Heys (1989) thought H. arumbera was constructed in the same manner as H. canadensis. Uchman (1995), however, interpreted Saerichnites as an interconnected, zigzag-branching burrow system, similar to Treptichnus. We disagree with the Uchman (1995) synonymy, and tentatively place H. canadensis and H. arumbera within Treptichnus vagans due to their beaded-chain morphology and the synonymization of Hormosiroidea under Halimedides Lorenz von Liburnau, 1902 (Uchman, 1998, 1999; Gaillard & Olivero, 2009).

  • Treptichnus vagans is also similar to “rosary-like structures” from the middle Cambrian Burgess Shale (Caron & others, 2010; Mángano, 2011) and linear Treptichnus bifurcus from the Lower Jurassic East Berlin Formation of Massachusetts (Getty & others, 2016). The Burgess Shale “rosary structures” are short to long, meandering to winding to zigzagging chains of small, beadlike mounds or pits with connecting tunnels and interpreted as chains of globular to spherical chambers used as agrichnia to farm bacteria (Mángano, 2011). Most Treptichnus vagans specimens lack tunnels connecting the beads. Some “rosary” chambers were filled with pyrite—which the T. vagans specimens lack—and were noted to support an agrichnial interpretation in dysoxic waters and nearanoxic sediments (Mángano, 2011). The “rosary structures” were also noted for their similarity to T. pollardi and its associated vertical shaft nodes (Jensen in Mángano, 2011) and are included in T. vagans. Specimens of linear T. bifurcus reported from the Lower Jurassic East Berlin Formation (Getty & others, 2016, fig. 4.5), described as “string of beads” and composed of linearly oriented burrow segments and swelled projection occurring end on end, are morphologically similar to the Burgess Shale rosary structures, Ptychoplasma, T. pollardi, and T. vagans. We, therefore, include the linear T. bifurcus specimens of Getty and others (2016) within T. vagans.

  • Figure 23.

    Treptichnus pedum and T. vagans “string of beads” specimens from the Spence Shale. 1, Treptichnus pedum in convex hyporelief, IBGS PJ-M-017, Miner's Hollow Float; 2, Treptichnus vagans with pitted furrow morphology, pits (arrows), in concave epirelief, IBGS LG-M-013, Antimony Canyon; 3, T. vagans with Planolites montanus (white arrow), Treptichnus-like P. montanus (white circle), and Rusophycus carbonarius (black arrow) ín convex hyporelief, KUMIP 314222 B, Miner's Hollow; 4, Eocene Treptichnus from the Green River Formation (Photo courtesy of Joshua Hogue, used with permission): (A) Concave epirelief, alternating beaded morphology (= beaded Treptichnus bifurcus); (B) transition to beaded single-chain morphology; and (C) pitted furrow morphology (= T. vagans); 5, Treptichnus vagans with Monomorphichnus bilinearis (arrows) in convex hyporelief, IBGS PJ-M-031, Miner's Hollow; scale bars in cm.

    f23_01.jpg

    Figure 24.

    Treptichnus vagans with variable bead morphology, including circular, fluted, imbricated, ovoid, and triangular morphologies from the Spence Shale. 1, Treptichnus vagans with showcased beads highlighted, KUMIP 314222 A, Miner's Hollow; 2, pair of smooth ovoid beads; 3, Circular (white arrow) and fluted (grey arrow) beads; 4, fluted circular bead (white arrow) and imbricated ovoid bead (black arrow); 5, circular and triangular beads (arrow); and 6, smooth ovoid beads; scale bar in (1) cm; scale bar (2–6) in mm.

    f24_01.jpg

    Miscellanea
    Jellyfish Impression?
    Figure 22.4–22.5

  • Material.—KUMIP 314121: one specimen (part and counterpart), Wellsville Mountains, Utah, USA.

  • Diagnosis.—Circular, convex mound (part) with broad, shallow depression near center.

  • Description.—Convex mound: 43.9–46.6 mm wide; 11.7 mm thick; depression is 2.8 mm deep. Elrathia? sp. trilobite mold on counterpart: 10.1 mm long, 7.2 mm wide; corresponds to dark gray ovoid-shaped area on part specimen.

  • Occurrence.—Gray, siliciclastic silty shale.

  • Associated ichnotaxa.—None.

  • Discussion.—The exact nature of this specimen is unknown but we propose several possible interpretations: (1) a body fossil and ichnofossil of an unknown cnidarian jellyfish (likely a scyphozoan) perhaps with and the Elrathia? sp. trilobite feeding off the remains of the jellyfish (i.e., Mortichnia and Praedichnia); (2) a resting trace of a suspension-feeding cnidarian, for instance, an upside-down jellyfish (Rhizostomeae, Cassiopeidae) or sea anemone; or (3) the nesting trace of an unknown tracemaker, similar to modern-day fish nests.

  • A cnidarian body fossil and ichnofossil interpretation is the most likely as there are reports of similar circular-shaped fossils interpreted as jellyfish body fossils (e.g., Hagadorn, Dott, & Damrov, 2002; Gaillard & others, 2006; Oosterink & Winkelhorst, 2013). Hagadorn, Dott, and Damrov (2002) and Gaillard and others (2006) illustrated jellyfish specimens with some of the internal morphology (e.g., gonads) and tentacles preserved, whereas other specimens only had concentric rings or slight deformation attributed to shrinkage and/or locomotive pulsation and localized downslope sliding, respectively. Oosterink and Winkelhorst (2013) specimens had concentric rings attributed to shrinkage and appeared to exhibit some internal morphology. The Spence Shale specimen lacks concentric rings to indicate pulsation, no deformation to indicate downslope sliding, or any discernible internal morphology. The central depression, however, may have been formed by the collapse of the jellyfish bell during decay. Trilobites are also known for being predators and/or scavengers of soft-bodied faunas (e.g., Jensen, 1990; Tarhan, Jensen, & Droser, 2011) and there is a report of a complex Rusophycus association that was interpreted as trilobites consuming possible jellyfish remains (Brandt & Rudkin, 2011). The close association of the jellyfish body impression and the Elrathia? sp. may represent predation or scavenging by the trilobite.

  • The second proposed interpretation of the mound is as a resting and/or suspension-feeding trace of an unknown species of upsidedown jellyfish (i.e., Cassiopeidae) or actinian (e.g., sea anemone). The upside-down jellyfish, Cassiopeia Perón & Lesueur, 1810, has a flat to concave, broad bell with tentacles extended upward to capture prey, and it commonly rests on the seafloor (Hummelinck, 1968; DeFelice, Eldridge, & Carlton, 2001; Schembri, Deidun, & Vella, 2010). The concave bell of Cassiopeia could possibly form a short, broad mound while resting on the sediment-water interface. An actinian tracemaker could also produce a similar shallow-mound form (e.g., Bergaueria sucta); however, the orientation of the specimen would be opposite of the current interpretation.

  • A third, proposed interpretation is that the mound is a nestlike excavation of an unknown tracemaker, possibly the Elrathia? sp. trilobite. Nestlike excavations of known and unknown tracemakers are not unheard of in ichnotaxonomy. Fenton and Fenton (1937d) established and interpreted Rusophycus jenningsi as a trilobite brooding nest. Ancient and modern fish produce simple to intricate, radially symmetric mounds or depressions (e.g., Piscichnus) to attract mates and spawn (e.g., Feibel, 1987; Hasiotis & others, 2012; Kawase, Okata, & Ito, 2013). The Elrathia? trilobite may have produced the mound in an attempt to attract a mate with whom to reproduce.

  • DISCUSSION

    Ichnotaxa

    Thirty-five ichnospecies were identified from 24 ichnogenera on slab specimens from the Spence Shale; Archaeonassa, Arenicolites, Aulichnites, Bergaueria, Conichnus, Cruziana, Dimorphichnus, Diplichnites, Gordia, Gyrophyllites, Halopoa, Lockeia, Monomorphichnus, Nereites, Phycodes, Phycosiphon, Planolites, Protovirgularia, Rusophycus, Sagittichnus, Scolicia, Taenidium, Teichichnus, and Treptichnus (Table 1).

    Behaviors

    Ichnofossils from the Spence Shale represent a variety of behaviors grouped into ethological categories (e.g., Bromley, 1996; Gingras & others, 2007) (Table 1); cubichnia (resting), domichnia (dwelling), fodinichnia (feeding), pascichnia (grazing), praedichnia (predation), and repichnia (locomotion). Cubichnia is represented by Lockeia, Rusophycus, and Sagittichnus. The ichnogenera of Arenicolites, Bergaueria, and Conichnus are commonly interpreted as domichnia. Traces commonly interpreted as fodinichnia include Gordia, Gyrophyllites, Halopoa, Phycodes, Planolites, Scolicia, Taenidium, Teichichnus, and Treptichnus. Pascichnia is represented by Cruziana, Gordia, Nereites, and Phycosiphon. Praedichnia are represented by compound ichnofossil associations of Rusophycus with Planolites and Archaeonassa jamisoni with Taenidium cf. satanassi, which represent epifaunal traces superimposed over the infaunal traces. Repichnia include Archaeonassa, Aulichnites, Cruziana, Dimorphichnus, Diplichnites, Monomorphichnus, and Protovirgularia.

    Ichnocoenoses

    An ichnocoenosis is an assemblage of ichnofossils that is the result of a single community of tracemaking organisms and can be used to interpret various physicochemical controls present during deposition (e.g., Ekdale, Bromley, & Pemberton, 1984; Bromley, 1996). Three ichnocoenoses are established for the Spence Shale, with a varying degree of stratigraphic occurrence: Rusophycus-Cruziana, Sagittichnus, and Arenicolites-Conichnus (Table 2). The ichnocoenoses suggest the Spence Shale was predominantly controlled by benthic oxygenation (Fig. 25–26).

    The Rusophycus-Cruziana (RC) ichnocoenosis occurs in gray to greenish gray, calcareous or siliciclastic silty shale. Four slab samples were assigned to the RC ichnocoenosis, but only two slabs (KUMIP 204523A+B and IBGS PJ-M-007) could be stratigraphically placed within the Spence Shale, and both occur near the base of Miner's Hollow Cycle 6 (see Fig. 5). The RC ichnocoenosis has the second highest ichnodiversity with seven ichnogenera present: Bergaueria, Cruziana, Lockeia, Monomorphichnus, Planolites, Rusophycus, and Treptichnus. The dominant behaviors represented are cubichnia, pascichnia, and repichnia with minor behaviors including fodinichnia, domichnia, and praedichnia. Ichnofabric indices range from ii1–2; whereas, bedding-plane bioturbation indices range from BPBI 2–4. The ichnocoenosis represents deposition in a proximal position on the outer detrital belt (see Fig. 2) (e.g., Robison, 1976, 1991; Liddell, Wright, & Brett, 1997) with; (1) low to moderate depositional energy; (2) low sedimentation rate; (3) low to moderate benthic oxygen but poorly oxygenated sediment; (4) moderate to high nutrients; and (5) minor bottom water currents (Fig. 26.1).

    The Sagittichnus ichnocoenosis is found in gray to greenish gray, siliciclastic silty shale with black dendrites and rarely interlaminated with calcareous shale. None of the assigned slab samples could be stratigraphically placed but are known from Antimony and Cataract canyons, Wellsville Mountains, and High Creek Canyon, Bear River Range. This ichnocoenosis has the highest ichnodiversity in the Spence Shale with eight ichnogenera represented: Bergaueria, Gyrophyllites, Lockeia, Planolites, Rusophycus, Sagittichnus, Teichichnus, and Treptichnus. The dominant behavior represented is cubichnia, and minor behaviors include fodinichnia and repichnia. The Sagittichnus ichnocoenosis represents deposition in a medial position on the outer detrital belt (e.g., Robison, 1976, 1991; Liddell, Wright, & Brett, 1997) (see, Fig. 2) with: (1) low to moderate depositional energy; (2) rapid sedimentation pulses with some tempestites; (3) low to moderate benthic oxygen; and (4) moderate nutrients (Fig. 26.2).

    Table 1.

    Frequency, preservation, behavioral ethologies, and trace makers of Spence Shale ichnofossils. Frequency key: A=abundant (> 20 specimens); C=common (6 to 20 specimens); R=rare (2 to 5 specimens); VR=very rare (1 specimen); P=part, CP=counterpart.

    t01_01.gif

    The Arenicolites-Conichnus (AC) ichnocoenosis is the most unique ichnocoenosis from the Spence Shale as it represents different dominant behaviors and an entirely different ichnofacies. The AC ichnocoenosis is from a float sample from Cataract Canyon and could not be stratigraphically placed. The dominant behaviors represented are domichnia and cubichnia. The AC ichnocoenosis also has a low ichnodiversity with only two ichnogenera represented: Arenicolites and Conichnus. The AC ichnocoenosis represents deposition in a proximal position near the boundary between the outer detrital belt and outer carbonate belt (e.g., Robison, 1976, 1991; Liddell, Wright, & Brett, 1997) (see Fig. 2) with: (1) moderate to high depositional energy; (2) moderate to high sedimentation; (3) moderate to high oxygen; and (4) medium (Fig. 26.3).

    Ichnofacies

    The majority of the ichnotaxa suggests that a significant portion of the Spence Shale was deposited in a distal Cruziana Ichnofacies. Bergaueria, Cruziana, Diplichnites, Monomorphichnus, and Rusophycus in the Rusophycus-Cruziana and Sagittichnus ichnocoenoses are the most indicative of this ichnofacies (Bromley, 1996; MacEachern & others, 2007a). The high number of pascichnial ichnofossils, small burrow diameters (e.g., Cruziana and Rusophycus), shallow sediment penetration, and low ii suggest that bottom-water oxygenation (likely dysoxia) influenced the biota and their behavior (e.g., Ekdale & Mason, 1988; MacEachern & others, 2007b; Gars on & others, 2012). Specimens assigned to the Cruziana Ichnofacies occur mostly in the silty shales near the base of Miner's Hollow Cycle 5 and 6 between 42–49 m above the Spence Shale base (see Fig. 5).

    Table 2.

    Ichnocoenoses of the Spence Shale with minor associated traces, dominant behaviors, and environmental interpretations.

    t02_01.gif

    The second ichnofacies proposed for the Spence Shale–based on a sample containing Arenicolites, Conichnus, ripple marks, and soft-sediment deformation—is a depauperate, distal Skolithos Ichnofacies indicating a higher energy environment with shifting media (MacEachern & others, 2007a, 2007b) (see Fig. 2). The depauperate, distal Skolithos Ichnofacies is present in peloidal carbonate wackestone to packs tone to mudstone and silty siliciclastic shale of the Spence Shale at the Cataract Canyon locality. The stratigraphic position of the Skolithos Ichnofacies is not known, as no stratigraphic data exists for the assigned sample.

    Comparative Ichnotaxonomy

    The Spence Shale ichnofauna is composed of numerous common facies-crossing ichnotaxa, which are represented in multiple depositional environments throughout the Phanerozoic. Similarities between the Spence Shale ichnotaxa and the ichnotaxa of other Cambrian-aged deposits suggest that shaley portions of the Spence Shale may have been deposited in shallow marine as well as deep settings following a fluctuating oxycline (sensu Garson & others, 2012).

    Ichnotaxonomy of BST Deposits.—Since there have been no ichnological studies of the Wheeler and Marjum formations, a comparison between Utah BST deposits is not possible. The Kaili Biota, Kaili Formation of China is the only other middle Cambrian BST deposit that has been extensively studied ichnologically. Other Cambrian BST deposits with reported ichnofossils include the early Cambrian Sirius Passet Biota, Buen Formation of Greenland and Chengjiang Biota, Yu'anshan Formation of China, and middle Cambrian Burgess Shale of British Columbia (Table 3).

    The Kaili Biota (Oryctocephalus indicus Biozone) (see Fig. 4.2) of the lower—middle Cambrian Kaili Formation in Guizhou Province, China, contains 26 ichnogenera (see Lin & others, 2010, appendix A, for complete list and references) and shares 10 ichnogenera in common with the Spence Shale (Table 3). Yang (1994) assigned the Kaili Biota to the Cruziana Ichnofacies and suggested that the Kaili Formation was deposited during near normal marine conditions in a shallow, nearshore setting under moderate to low energy. Lin and others (2010) suggested that the major sedimentation events of the Kaili Formation occurred due to episodic distal tempestites with relatively low background sedimentation. The Kaili Formation distal tempestite deposition is similar to the Robison (1991) suggestion that many of the Spence Shale Fagerstätten were deposited by tempestites in the distal ramp setting of the Spence Shale. Sudden burial by tempestites (i.e., obrusion) may produce anoxic—dysoxic conditions in the underlying sediment, enabling the production of BST fossils until oxic conditions returned, allowing organisms to burrow, mix sediments, and even feed on the preserved soft tissues (Garson & others, 2012).

    The Sirius Passet Biota (SPB) from the early Cambrian (Series 2, Stage 3) (see Fig. 4.1) of Greenland is a remote but rich BST deposit with only six ichnogenera, sharing three with the Spence Shale (Ineson & Peel, 2011). Most of the traces reported from the Sirius Passet were simple, horizontal meandering burrows (likely Gordia, Helminthoidichnites, and Planolites based on photographs) with some specimens of Chondrites, Cosmorhaphe?, Megagrapton?, Palaeophycus, Planolites, Spirorhaphe?, and Teichichnus (Table 3) (Ineson & Peel, 2011). Mángano and others (2012) examined narrow, filamentlike structures similar to Pilichnus yet no formal assignment was made; however, the SPB “Pilichnus” is more likely to be a tubular body fossil similar to Vendotaenia antiqua (e.g., Cohen & others, 2009), which is considered analogous to a green or red alga. The SPB was deposited in the deep-water shales of the Buen Formation as part of an outer shelf and slope environment (Peel, 2010). Pyrite is present in the burrow fill of some SPB ichnofossils, suggesting an oxygen-depleted environment (sensu Martin, 2004; Ineson & Peel, 2011). No ichnofacies has been assigned to the Buen Formation but likely contains a Zoophycos or Nereites Ichnofacies.

    The Chengjiang Biota of the Yu'anshan Formation of the early Cambrian of Yunnan Province, China, has had several reports of ichnofossils in close association with BST fossils (e.g., Zhang & others, 2007; Huang & others, 2014) (Table 3). Zhang and others (2007) reported small (< 2.0 mm diameter), unidentified ichnofossils that burrowed through and beneath BST films, similar to Gordia specimens in Wang and others (2009), and suggested they may be forms of Helminthoidichnites or Pilichnus. Huang and others (2014) had several worm specimens interpreted to have died within thinly lined, horizontal to subvertical burrows—some of which were reported as U shaped with paired openings—but no ichnogeneric names were assigned. These morphologies could represent specimens of Arenicolites, Palaeophycus, Planolites, or Skolithos. The Chengjiang ichnofossils suggest a distal Cruziana Ichnofacies.

    Figure 25.

    Primary physicochemical controls on organism behavior. 1, Established primary physicochemical controls in marine depositional systems; 2, interpreted physicochemical controls for the Spence Shale (modified from Hasiotis & Platt, 2012).

    f25_01.jpg

    The Burgess Shale of the middle Cambrian of British Columbia has had few ichnofossils reported—most reported in open nomenclature (e.g., “[U]-shaped tube trace” and “vertical-pipe morphology”) by Allison and Brett (1995, fig. 4)—and shares five of eleven ichnogenera with the Spence Shale (Table 3). Hagadorn (2002) assigned the Allison and Brett (1995) ichnofossils to Arenicolites, Cruziana, Monocraterion, and Planolites; however, the U-shaped tubes were also described as having reworked sediment between the arms, which would place them in Diplocraterion. Caron and others (2010) reported the first ichnotaxonomically assigned ichnofossils from the Burgess Shale (as the “thin” Stephen Formation), including Cruziana problematica, Diplichnites, Gordia, Helminthoidichnites, and a pellet-filled burrow, Alcynidiopsis, filled with coprolites (possibly Tibikoia or Tomaculum) associated with an arthropod carapace. These ichnofossils, however, are only illustrated in the supplementary materials (see Caron & others, 2010, supplementary material 8 GSA Data Repository 2010228). Mángano (2011) reexamined the material of Caron and others (2010) and reported specimens of Diplopodichnus and Helminthopsis. Several large arthropod trackway sets were described from the Kicking Horse Member (Glossopleura biozone) of the Burgess Shale Formation as Diplichnites (Minter, Mángano, & Caron 2012). Cheiichnus, Fuersichnus, and arthropod trackway specimens were reported from near the base of the Stephen Formation (Caron & others, 2014, supplementary fig. 3–5). The Fuersichnus specimens are more likely specimens of Palaeophycus or Phycodes due to similar morphologies and lack of retrusive spreiten (e.g., Bromley & Asgaard, 1979; Ekdale, Bromley, & Pemberton, 1984; Hasiotis, 2002; Garvey & Hasiotis, 2008). Mángano (2011) interpreted media consistency (substrate control) and benthic oxygenation as the primary physicochemical controls on the Burgess Shale ichnofauna. The Burgess Shale ichnofauna likely represent shifts between a distal Cruziana and Zoophycos ichnofacies.

    Ichnotaxonomy of Non-BST Cambrian deposits.—The Spence Shale shares ichnotaxa with multiple non-BST-bearing Cambrian deposits (Table 4).

    The Cándana Quartzite of the Ediacaran—early Cambrian of northern Spain has reported 18 ichnogenera (Crimes & others, 1977) and shares 11 ichnogenera in common with the Spence Shale (Table 4). The Cándana Quartzite was deposited in tidal channels, and intertidal and sub tidal sand bars (Crimes & others, 1977). No ichnofacies was assigned, but the ichnofossils present suggest a Cruziana Ichnofacies.

    The Chapel Island and Random formations of the Ediacaran— early Cambrian in Canada has 27 ichnogenera (e.g., Crimes & Anderson, 1985; Droser & others, 2002) with 11 ichnogenera in common with Spence Shale (Table 4). The Cambrian-aged sections of the Chapel Island and Random formations record a transition from an offshore to prograding delta front to tidal-channel and tidal-flat setting. No ichnofacies was assigned to either the Chapel Island or Random formation, but likely contains a shift from a Cruziana Ichnofacies to Skolithos Ichnofacies. Most of the ichnogenera shared with the Spence Shale occur in the upward-thickening siltstones, mudstones, and thinly bedded sandstones of the prograding delta front, shoreface rippled siltstones and sandstone, or shifting sand bars and channels.

    The Arumbera Sandstone of the Ediacaran—early Cambrian of central Australia contains 24 ichnogenera and shares 11 in common with the Spence Shale (Wells & others, 1970; Walter, Elphinstone, & Heys, 1989) (Table 4). The Cambrian-aged upper half of the Arumbera Sandstone—which contains the majority of the ichnotaxa—was deposited in a shallowing, marine basinal to shoreface to prograding coastal delta-plain setting. While no ichnofacies was assigned, the Arumbera Sandstone likely contains two ichnofacies, Cruziana and Skolithos ichnofacies, and possibly a third, Nereites Ichnofacies. The Cruziana (and possible Nereites) Ichnofacies likely occurs in the gray-green shales interbedded with thin sandstones interpreted as basinal deposits. The Skolithos Ichnofacies likely occurs in the thick sandstones of the shoreface and prograding delta-plain deposits.

    Figure 26.

    Spence Shale ichnocoenosis models and interpreted physicochemical controls. 1, Rusophycus-Cruziana ichnocoenosis, dominant control: benthic oxygenation; 2, Sagittichnus ichnocoenosis, dominant control: benthic oxygenation; 3, Arenicolites-Conichnus ichnocoenosis, dominant control: depositional energy.

    f26_01.jpg

    Table 3.

    Ages, depositional environments, ichnofacies, and shared ichnotaxa of Cambrian BST deposits.

    t03_01.gif

    The Holy Cross Group (HCG) of the early Cambrian—Early Ordovician of the Polish Holy Cross Mountains contains nine formations ranging from shallow to deep marine and shares 16 of 43 ichnogenera with the Spence Shale (e.g., Orłowski, 1989, 1992; Orłowski & Żylińska, 2002; Stachacz, 2016) (Table 4). Six formations of the HCG were deposited during the middle Cambrian, but most units had low ichnodiversity (1–5 ichnogenera) except the early—middle Cambrian Ociesęki Sandstone Formation with a high ichnodiversity (43 ichnogenera; e.g., Orłowski, 1989, 1992; Orłowski & Żylińska, 2002; Stachacz, 2016). Middle Cambrian HCG units are composed mostly of clayey to silty shales and siltstones intercalated in fine-grained sandstones (Orłowski, 1989). The majority of ichnofossils from the HCG were assigned to the Cruziana Ichnofacies (e.g., Orłowski, 1989, 1992; Orłowski & Zylińska, 2002; Stachacz, 2012), whereas some specimens are representative of the Nereites Ichnofacies (Orłowski & Żylińska, 2002) and the Skolithos Ichnofacies in the upper portions Ociesęki Sandstone Formation (Stachacz, 2016).

    The Mickwitzia Sandstone Member of the File Haidar Formation of the early Cambrian in Sweden is a shallow-marine unit deposited over a Precambrian basement and shares 8 of 24 ichnogenera with the Spence Shale (Table 4). The Mickwitzia Sandstone is composed mostly of thin-bedded, fine- to coarsegrained sandstones and siltstones interbedded with claystone, and a conglomeritic base. The majority of Mickwitzia ichnofossils (e.g., Cruziana, Gyrolithes, Rosselia, Rusophycus, and Zoophycos) occur in thinly bedded sandstone and siltstone on a mud-dominated shallow shelf assigned to the Cruziana Ichnofacies. Some intervals were assigned to the Skolithos Ichnofacies. Intervals assigned to the Cruziana Ichnofacies typically had an ii2–3; whereas, intervals assigned to the Skolithos ichnofacies had an ii3–4 (Jensen, 1997).

    The Paseky Shale of the early Cambrian of the Czech Republic is a restricted shallow-marine, brackish lagoon or estuary, and shares all five ichnogenera with the Spence Shale (Table 4). The Paseky Shale is composed of alternating claystone and siltstone with fine-grained graywacke intercalations and numerous adhesion structures and wrinkle marks (Kukal, 1995). Paseky ichnofossils are restricted to a 3-m-thick section of light green, olive-, or graygreen laminated shale (Mikuláš, 1995). Most marine ichnotaxa are missing from the Paseky Shale indicating a continental or restricted marine environment (Mikuláš, 1995). Though not discussed by Mikuláš (1995), the reported ichnotaxa are suggestive of the Cruziana Ichnofacies.

    The lower Cambrian (Terreneuvian—Series 2) of the White-Inyo Mountains, eastern California, USA, consists of five formations (Deep Spring, Campito, Poleta, Harkless, Saline Valley, and Mule Spring formations) of alternating terrigenous-clastic and carbonate sandstones and shales deposited on a shallow, subtidal shelf (e.g., Marenco & Bottjer, 2008). The White-Inyo Mountains contain 28 ichnogenera with 11 ichnogenera in common with the Spence Shale (e.g., Alpert, 1973, 1976a, 1976b; Alpert & Moore, 1975; Marenco & Bottjer, 2008) (Table 4). The majority of ichnofossils occur in micaceous siltstone and cross-bedded sandstones. The Alpert (1976a, 1976b) ichnofossils suggest multiple ichnofacies are recorded in the White-Inyo Mountains; (1) the Deep Spring Formation likely contains a distal Skolithos Ichnofacies due to the presence of Diplichnites, Monocraterion (rare), Monomorphichnus, Planolites (common), Rusophycus, and Skolithos (rare); (2) the Campito Formation likely records a shift from a distal Cruziana to proximal Cruziana Ichnofacies due to a shift in the ichnofossil suite from Archaeonassa, Belorhaphe, Bergaueria, Cochlichnus, Helminthopsis, Rusophycus, and Scolicia in the Andrews Mountain Member to Archaeonassa, Astropolithon?, Dactyloidites, Monocraterion, Planolites, Skolithos, and Teichichnus in the Montenegro Member; (3) the Poleta Formation likely contains a distal Skolithos Ichnofacies due to the presence of Archaeonassa, Arthrophycus?, Bergaueria, Dolopichnus, Laevicyclus, Monocraterion, Planolites, Psammichnites, Rusophycus, Scolicia, Skolithos (common), and Teichichnus; and (4) the Harkless Formation likely contains an archetypal Cruziana Ichnofacies due to the presence of Archaeonassa, Asteriacites?, Bergaueria, Cruziana, Diplichnites, Monocraterion, Monomorphichnus, Planolites, Rusophycus, Scolicia, Skolithos, and Teichichnus. Alpert (1976a, 1976b) reported only a few ichnofossils from the Saline Valley (i.e., Cruziana, Planolites, and Teichichnus), and did not mention any from the Mule Spring Formation. Mount (1982) later assigned the Andrews Mountain Member of the Campito Formation to the Cruziana Ichnofacies.

    Table 4.

    Ages, depositional environments, ichnofacies, and shared ichnotaxa of Cambrian non-BST deposits.

    t04_01.gif

    The Bright Angel Shale (BAS) of the Grand Canyon area was deposited approximately at the same time as the Spence Shale (Cambrian, Series 3), and has been assigned to the Glossopleura trilobite biozone (Baldwin & others, 2004). The age and location of the BAS places it within the inner detrital belt of Robison (1960; see Fig. 2). The BAS shares 11 of 21 ichnogenera with the Spence Shale (Table 4). Low energy, silty and muddy laminated beds of the BAS dominated by Cruziana and Diplichnites are similar to Spence Shale beds containing C. problematica. There is still some debate, however, concerning the depositional environment of the BAS. Elliot and Martin (1987) and Lane and others (2003) proposed the BAS was deposited in a shelf environment influenced by both tides and storms; whereas, Baldwin and others (2004) argued the BAS is an estuarine deposit due to high concentrations of freshwater palynomorphs in the heterolithic sandstones and shales. While no ichnofacies was formally assigned to the BAS, Baldwin and others (2004) noted that elements of the Skolithos and Cruziana ichnofacies tend to mix and are juxtaposed within the same beds and could be assigned to a mixed Skolithos-Cruziana ichnofacies.

    The Hanneh Member of the Burj Formation of the middle Cambrian in the Dead Sea Basin, Jordan, contains 19 ichnogenera and was deposited in a shallow marine prodelta—delta-front to tidal-flat system (Hofmann & others, 2012; Mangano & others, 2013). The Hanneh Member is composed mostly of siliciclastic mudstone and crossbedded to laminated, fine- to medium-grained sandstone with a limestone base, with ichnofossils present in mudstone and sandstone. Twelve of the 19 Hanneh ichnogenera are also present in the Spence Shale (Table 4). Two ichnofacies were assigned to the Hanneh Member: a Glossifungites Ichnofacies represented by Diplocraterion—suggesting high depositional energy and significant erosion—and a Cruziana Ichnofacies represented by Cruziana, Diplichnites, Gyrolithes, and Rusophycus—suggesting soft to firmground media in low-energy settings.

    The upper Cambrian?—Lower Ordovician Bell Island and Wabana groups of Newfoundland, Canada, is well-studied shallow marine (e.g., onshore lagoon, tidal flat, delta front, etc.) to offshore transition zone deposits. The Bell Island and Wabana groups contain 39 ichnogenera and share 17 ichnogenera with the Spence Shale (Table 4). Many of the shared ichnofossils occur in delta front, middle tidal flat, and lagoonal deposits. The domichnia-dominated sandstones of the subtidal, shoreface, foreshore, and sandbar deposits were assigned to the Skolithos Ichnofacies; whereas the fodinichnia-dominated thin sandstones and shales of the intertidal-flat, lagoonal, and delta-front deposits were assigned to the Cruziana Ichnofacies (Fillion & Pickerill, 1990).

    Summary.—The Spence Shale shares more ichnogenera in common with shallow-marine BST deposits than with most deep-marine BST deposits. The moderate number of domichnia and repichnia ichnofossils present and higher ichnodiversity of the Spence Shale is more similar to shallow-marine BST deposits (i.e., Kaili Biota), suggesting that the ichnofossil-bearing beds of the Spence Shale were deposited in a shallower environment on the distal ramp than previously thought. Low ii(ii2–3) but low to high BPBI (BPBI 2–5) suggest the Spence Shale ichnofauna was predominantly controlled by fluctuating benthic oxygenation (sensu Garson & others, 2012). Periodic tempestite deposition (Robison, 1991) and soft-sediment deformation, orientated Rusophycus and ripple marks, and frequent pascichnia suggest sedimentation rate, depositional energy, and nutrient availability also had significant influence on the Spence Shale ichnofauna, respectively. The degree of similarity between the Spence Shale and Kaili Biota ichnofaunas, however, may also be to due to the fact that deep-marine BST deposits are more understudied ichnotaxonomically than their shallow-marine counterpart. When more ichnotaxonomic research on deep-marine BST deposits is available, a better comparison can be made.

    The Spence Shale ichnofauna is similar to both shallow- and deep-marine, non-BST Cambrian ichnofaunas. The Spence Shale ichnofauna occurs in calcareous to siliciclastic, silty to sandy shales to sandstone like the BAS, Chapel Island and Random, HCG, Mickwitzia Sandstone, and Paseky Shale ichnofaunas. The similarity between the non-BST shallow- and deep-marine ichnofauna and lithofacies associations suggests the Spence Shale was deposited, in part, on a middle section of the distal ramp, where controlling physicochemical factors from shallow and deep settings (e.g., benthic oxygenation, depositional energy, and nutrient availability) could influence the Spence Shale ichnofauna (see Fig. 2). The lack of extensive endobenthic fodinichnia (e.g., Chondrites, Cosmorhaphe, or Zoophycos) or agrichnia (e.g., Paleodictyon)—suggestive of low energy and high dysoxia—suggests the studied shales of the Spence Shale were not basinal deposits. The lack of extensive or reinforced domichnia (e.g., Palaeophycus, Skolithos, or Thalassinoides) or reworked equilibrichnia (e.g., Diplocraterion)—suggestive of high depositional energy and/or rapidly shifting media—suggests the studied shales were not deposited proximally to the carbonate platform.

    CONCLUSIONS

    1. The Spence Shale is most known for its numerous and highly well-preserved body fossils, especially trilobites. The Spence Shale is now also known for its numerous ichnofossils and highly diverse ichnofossil assemblage with 24 ichnogenera and 35 ichnospecies. Ichnofossils of the Spence Shale primarily occur in light to dark gray, calcareous or siliciclastic shale, and represent cubichnia, domichnia, fodinichnia, pascichnia, praedichnia, and repichnia behaviors.

    2. A new ichnospecies, Archaeonassa jamisoni, is proposed for short, downward excavations with rimmed margins. Ptychoplasma (Protovirgularia) vagans is emended and transferred to Treptichnus as T. vagans.

    3. Three ichnocoenoses were constructed and two ichnofacies were assigned to the Spence Shale: (1) a distal Cruziana Ichnofacies representing low- to moderate-energy deposition in oxygen- and nutrient-controlled ichnocoenoses (e.g., Rusophycus-Cruziana and Sagittichnus); and (2) a depauperate, distal Skolithos Ichnofacies representing moderate- to high-energy deposition with Arenicolites and Conichnus as representative ichnotaxa (i.e., Arenicolites-Conichnus ichnocoenosis). The Spence Shale ichnofauna was controlled by benthic oxygenation, depositional energy, and nutrient availability.

    4. The Spence Shale contains numerous BST fossils and ichnofossils and has the second highest known ichnodiversity of BST deposits, and shares more ichnotaxa in common with shallow-marine systems (∼11–12 ichnogenera; e.g., Kaili Biota, Hanneh Member of the Burj Formation) than deep-marine systems (∼2–4 ichnogenera; e.g., Burgess Shale, Sirius Passet Biota of the Buen Formation), suggesting deposition on shallower parts of the distal ramp setting.

    ACKNOWLEDGEMENTS

    We thank everyone who helped and supported this paper: the Gunther Family (Lloyd and Val Gunther), Paul Jamison, and Phillip Reese for collecting and donating much of the specimens; Dr. Richard A. Robison for introducing us to this project and to the Gunther Family and Paul Jamison, collecting and donating specimens, and for discussions concerning the history of the Spence Shale and the middle Cambrian of North America; the IchnoBioGeoScience (IBGS) Research Group at the University of Kansas (KU) for editing and discussing this paper and several conference presentations, especially Joshua Hogue for discussions on his Eocene treptichnid specimens; Dr. Sören Jensen for his input and help in identifying some of the ichnofossils; Dr. Jean-Bernard Caron for his input and identification of the Burgess Shale-type fossil presented herein; Dr. Sören Jensen and an anonymous reviewer, as well as associate editor Dr. Mary Droser, for their comments and suggestions, which improved the manuscript; Dr. Roscoe Jackson II for providing funding with the 2015 Roscoe G. Jackson II Graduate Research Award; KU Geology Associates Program for the 2014 Raymond C. and Lilian B. Moore Scholarship; and Daniel F. and Annie L. Merriam for the 2013 Merriam Graduate Student Research Award.

    DEDICATION

    This paper is dedicated to the memory of Lloyd Gunther (1917–2013), the patriarch of the generous fossil-collecting Gunther family, whose large contributions and donations of fossil specimens to numerous museums and universities have enabled a significant portion of Earth's past biodiversity to come alive again.

    REFERENCES

    1.

    Allison, P. A., & C. E. Brett. 1995. In situ benthos and paleo-oxygenation in the Middle Cambrian Burgess Shale, British Columbia, Canada: Geology 23(12):1079–1082. Google Scholar

    2.

    Alonso—Mumaga, P. J., L. A. Buatois, & C. O. Limarino. 2013. Ichnology of the Late Carboniferous Hoyada Verde Formation of western Argentina: Exploring postglacial shallow-marine ecosystems of Gondwana: Palaeogeography, Palaeoclimatology, Palaeoecology 369:228–238. Google Scholar

    3.

    Alpert, S. P. 1973. Bergaueria Prantl (Cambrian and Ordovician), a probable actinian trace fossil: Journal of Paleontology 47(5) :919–924. Google Scholar

    4.

    Alpert, S. P. 1976a. Trilobite and star-like trace fossils the White-Inyo Mountains, California: Journal of Paleontology 50(2):226–239. Google Scholar

    5.

    Alpert, S. P. 1976b. Trace fossils of the White-Inyo Mountains. In J. N. Moore, & A. E. Fritsche, eds., Depositional Environments of Lower Paleozoic Rocks in the White-Inyo Mountains, Inyo County, California. Pacific Coast Paleogeography Field Guides. SEPM Pacific Section. Los Angeles, p. 43–48. Google Scholar

    6.

    Alpert, S. P., & J. N. Moore. 1975. Lower Cambrian trace fossil evidence for predation on trilobkes: Lethaia 8(3):223–230. Google Scholar

    7.

    Archer, A. W., & C. G. Maples. 1984. Trace-fossil distribution across a marine-to-nonmarine gradient in the Pennsylvanian of southwestern Indiana: Journal of Paleontology 58(2):448–466. Google Scholar

    8.

    Ash, S. R., & S. T. Hasiotis. 2013. New occurrences of the controversial Late Triassic plant fossil Sanmiguelia Brown and associated ichnofossils in the Chinle Formation of Arizona and Utah, USA: Neues Jahrbuch für Geologie und Paläontologie-Abhandlungen 268(1):65–82. Google Scholar

    9.

    Babcock, L. E., & R. A. Robison. 1988. Taxonomy and paleobiology of some Middle Cambrian Scenella (Cnidaria) and hyolithids (Mollusca) from western North America: The University of Kansas Paleontological Contributions 121:1–22. Google Scholar

    10.

    Badve, R. M., & M. A. Ghare. 1978. Jurassic ichnofauna of Kutch—I: Biovigyanam 4:125–140. Google Scholar

    11.

    Badve, R. M., & M. A. Ghare. 1980. Ichnofauna of the Bagh Beds from the Deva River Valley South of Narmada: Biovigyanam 6:121–130. Google Scholar

    12.

    Baldwin, C. T., P. K. Strother, J. H. Beck, & E. Rose. 2004. Palaeoecology of the Bright Angel Shale in the eastern Grand Canyon, Arizona, USA, incorporating sedimentological, ichnological and palynological data. In D. Mcllroy, ed., The Application of Ichnology to Palaeoenvironmental and Stratigraphic Analysis. Geological Society. London. Special Publication 228. p. 213–236. Google Scholar

    13.

    Bandel, K. 1967. Trace fossils from two Upper Pennsylvanian sandstones in Kansas: The University of Kansas Paleontological Contributions 18:1–13. Google Scholar

    14.

    Bandel, K. 1973. A new name for the ichnogenus Cylindrichnus Bandel, 1967: Journal of Paleontology 47(5):1002. Google Scholar

    15.

    Bather, F. A. 1925. U-shaped Burrows Near Blea Wyke: Yorkshire Geological Society Proceedings 20(2):185–199. Google Scholar

    16.

    Bednarczyk, W., & T. Przybyłowicz. 1980. On the development of Middle Cambrian sediments in the Gdańsk Bay area: Acta Geologica Polonica 30(4):391–415, 24 pl. Google Scholar

    17.

    Benton, M. J. 1982. Trace fossils from Lower Palaeozoic ocean-floor sediments of the Southern Uplands of Scotland: Transactions of the Royal Society of Edinburgh Earth Sciences 73(2):67–87. Google Scholar

    18.

    Bertling, M., S. J. Braddy, R. G. Bromley, G. R. Demathieu, J. Genise, R. Mikuláš, J. K. Nielsen, K. S. S. Nielsen, A. K. Rindsberg, & M. Schlirf. 2006. Names for trace fossils: a uniform approach: Lethaia 39(3):265–286. Google Scholar

    19.

    Billings, E. 1862. On some new species of fossils from different parts of the Lower, Middle, and Upper Silurian rocks of Canada, Palaeozoic fossils, Volume 1:1861–1865. Geological Survey of Canada Advance Sheets, p. 96–168. Google Scholar

    20.

    Billing, E. 1866. Catalogue of the Silurian fossils of the Island of Anticosti , with descriptions of some new genera and species: Canadian Geological Survey, 426:93 p., 28 text-fig. Google Scholar

    21.

    Binney, E. W. 1852. On some trails and holes found in rocks of the Carboniferous strata, with remarks on the Microconchus carbonarius: Memoirs and proceedings of the Manchester Literary and Philosophical Society 2(10):181–201. Google Scholar

    22.

    Bohacs, K. M., S. T. Hasiotis, & T. M. Demko. 2007. Continental ichnofossils of the Green River and Wasatch Formations, Eocene, Wyoming: a preliminary survey, proposed relation to lake-basin type, and application to integrated paleoenvironmental interpretation: The Mountain Geologist 44(2):79–108. Google Scholar

    23.

    Bornemann, J. G. 1889. Über den Buntsandstein in Deutschland und seine Bedeutung für die Trias, nebst Unterscuchungen über Sandund Sandsteinbildungen im Allgemeinen: In J. G. Bornemann, ed., Beiträge zur Geologie und Paläontologie 1:1–61. [In German]. Google Scholar

    24.

    Brady, L. F. 1947. Invertebrate tracks from the Coconino sandstone of Northern Arizona: Journal of Paleontology 2(5):466–472, text-fig. 1–2, pl. 66–69. Google Scholar

    25.

    Bradshaw, M. A. 1981. Paleoenvkonmental interpretations and systematics of Devonian trace fossils from the Taylor Group (lower Beacon Supergroup), Antarctica: New Zealand Journal of Geology and Geophysics 24(5–6) :615–652. Google Scholar

    26.

    Brandt, D. S., & D. Rudkin. 2011. A Curious Rusophycus (Arthropod Ichnofossil) Assemblage from the Upper Ordovician of Ontario, Canada: Ichnos 18(1):35–40. Google Scholar

    27.

    Briggs, D. E. G., B. S. Lieberman, J. R. Hendricks, S. L. Halgedahl, & R. D. Jarrard. 2008. Middle Cambrian Arthropods from Utah: Journal of Paleontology 82(2):238–254. Google Scholar

    28.

    Briggs, D. E. G., & R. A. Robison. 1984. Exceptionally preserved nontrilobite arthropods and Anomalocaris from the Middle Cambrian of Utah: The University of Kansas Paleontological Contributions 111:1–24. Google Scholar

    29.

    Briggs, D. E. G., W. D. I. Rolfe, & J. Brannan. 1979. A giant myriapod trail from the Namurian of Arran, Scotland: Palaeontolog, 22(2):273–291. Google Scholar

    30.

    Bromley, R. G. 1996. Trace fossils: Biology, Taphonomy and Applications. Chapman & Hall. London. 361 pp. Google Scholar

    31.

    Bromley, R., & U. Asgaard. 1979. Triassic freshwater ichnocoenoses from Carlsberg Fjord, east Greenland: Palaeogeography, Palaeoclimatology, Palaeoecology 28:39–80. Google Scholar

    32.

    Buatois, L. A., J. Almond, & G. J. Germs. 2013. Environmental tolerance and range offset of Treptichnus pedum: Implications for the recognition of the Ediacaran-Cambrian boundary: Geology 41(4):519–522. Google Scholar

    33.

    Buatois, L. A., & M. G. Mángano. 1993a. The ichnotaxonomic status of Plangtichnus and Treptichnus: Ichnos 2(3):217–224. Google Scholar

    34.

    Buatois, L. A., & M. G. Mángano. 1993b. Trace fossils from a Carboniferous turbiditic lake: implications for the recognition of additional nonmarine ichnofacies: Ichnos 2(3):237–258. Google Scholar

    35.

    Buatois, L. A., & M. G. Mángano. 2002. Trace fossils from Carboniferous floodplain deposits in western Argentina: implications for ichnofacies models of continental environments: Palaeogeography, Palaeoclimatology, Palaeoecology 183(1):71–86. Google Scholar

    36.

    Buatois, L. A., & M. G. Mángano. 2007. Invertebrate ichnology of continental freshwater environments. In W. Miller III, ed., Trace fossils: concepts, problems, prospects. Elsevier. Amsterdam, p. 285–323. Google Scholar

    37.

    Buatois, L. A., & M. G. Mángano. 2011. Ichnology: Organism-substrate interactions in space and time. Cambridge University Press. Cambridge. 366 p. Google Scholar

    38.

    Buatois, L. A., M. G. Mángano, C. G. Maples, & W. P. Lanier. 1998. Taxonomic reassessment of the ichnogenus Beaconichnus and additional examples from the Carboniferous of Kansas, U.S.A: Ichnos 5(4):287–302. Google Scholar

    39.

    Buatois, L. A., M. G. Mângano, X. Wu, & G. Zhang. 1996. Trace fossils from Jurassic lacustrine turbidites of the Anyao Formation (central China) and their environmental and evolutionary significance: Ichnos 4(4):287–303. Google Scholar

    40.

    Buatois, L. A., M. G. Mángano, W. Xiantao, & Z. Guocheng. 1995. Vagorichnus, a new ichnogenus for feeding burrow systems and its occurrence as discrete and compound ichnotaxa in Jurassic lacustrine turbidites of Central China: Ichnos 3(4):265–272. Google Scholar

    41.

    Buatois, L. A., C. E. Uba, M. G. Mángano, C. Hulka, & C. Heubeck. 2007. Deep and intense bioturbation in continental environments: Evidence from Miocene fluvial deposits of Bolivia. SEPM. Tulsa, Oklahoma. Special Publication 88. p. 123–136. Google Scholar

    42.

    Buchanan, J. B., & R. H. Hedley. 1960. A contribution to the biology of Astrorhiza limicola (Foraminifera): Journal of the Marine Biological Association of the United Kingdom 39(3):549–560. Google Scholar

    43.

    Buckman, J. O. 1994. Archaeonassa Fenton and Fenton 1937 reviewed: Ichnos 3(3):185–192. Google Scholar

    44.

    Butterfield, N. J. 1990. Organic Preservation of Non-Mineralizing Organisms and the Taphonomy of the Burgess Shale: Paleobiology 16(3):272–286. Google Scholar

    45.

    Butterfield, N. J. 1995. Secular distribution of Burgess Shale-type preservation: Lethaia 28:1–13. Google Scholar

    46.

    Callow, R. H., D. Mcllroy, & M. D. Brasier. 2011. John Salter and the Ediacara fauna of the Longmyndian Supergroup: Ichnos 18(3):176–187. Google Scholar

    47.

    Carmona, N. B., M. G. Mángano, L. A. Buatois, & J. J. Ponce. 2010. Taphonomy and paleoecology of the bivalve trace fossil Protovirgularia in deltaic heterolithic facies of the Miocene Chenque formation, Patagonia, Argentina: Journal of Paleontology 84(4):730–738. Google Scholar

    48.

    Caron, J.-B., R. R. Gaines, C. Aria, M. G. Mángano, & M. Streng. 2014. A new phyllopod bed-like assemblage from the Burgess Shale of the Canadian Rockies: Nature communications 5(3210):1–6. Google Scholar

    49.

    Caron, J.-B., R. R. Gaines, M. G. Mángano, M. Streng, & A. C. Daley. 2010. A new Burgess Shale—type assemblage from the “thin” Stephen Formation of the southern Canadian Rockies: Geology 38(9):811–814. Supplementary materials 1–8 (GSA Data Repository 2010228). Google Scholar

    50.

    Chamberlain, C. K. 1971. Morphology and ethology of trace fossils from the Ouachita Mountains, southeast Oklahoma: Journal of Paleontology 45(2):212–246. Google Scholar

    51.

    Chamberlain, C. K. 1975. Trace Fossils in DSDP Cores of the Pacific: Journal of Paleontology 49(6):1074–1096. Google Scholar

    52.

    Chamberlain, C. K. 1977. Ordovician and Devonian trace fossils from Nevada. Nevada Bureau of Mines & Geology. Bulletin 90:24 p. Google Scholar

    53.

    Chamberlain, C. K., & D. L. Clark. 1973. Trace Fossils and Conodonts as Evidence for Deep-Water Deposits in the Oquirrh Basin of Central Utah: Journal of Paleontology 47(4):663–682. Google Scholar

    54.

    Chiplonkar, G. W., & R. M. Badve. 1970. Trace fossils from the Bagh Beds—Part II: Journal of the Palaeontological Society of India 15:1–5. Google Scholar

    55.

    Chisholm, J. I. 1970. Teichichnus and related trace-fossils in the Lower Carboniferous at St. Monance, Scotland: Geological Survey of Great Britain Bulletin 32:21–51. Google Scholar

    56.

    Chisholm, J. I. 1985. Xiphosurid burrows from the lower coal measures (Westphalian A) of West Yorkshire: Palaeontology 28(4):619–628. Google Scholar

    57.

    Cohen, P. A., A. Bradley, A. H. Knoll, J. P. Grotzinger, S. Jensen, J. Abelson, K. Hand, G. Love, J. Metz, N. McLoughlin, P. Meister, R. Shepard, M. Tice, & J. P. Wilson. 2009. Tubular compression fossils from the Ediacaran Nama Group, Namibia: Journal of Paleontology 83(1):110–122. Google Scholar

    58.

    Collom, C. J., P. A. Johnston, & W. G. Powell. 2009. Reinterpretation of ‘Middle’ Cambrian stratigraphy of the rifted western Laurentian margin: Burgess Shale Formation and contiguous units (Sauk II megasequence), Rocky Mountains, Canada: Palaeogeography, Palaeoclimatology, Palaeoecology 277(1–2):63–85. Google Scholar

    59.

    Colorado Plateau Geosystems. 2007. Deep Time Maps: Maps of Ancient Earth—North American Regional Geology: Middle Cambrian (510 Ma), updated May 2016,  https://deeptimemaps.com. Checked May 2018. Google Scholar

    60.

    Conway Morris, S. 1986. The community structure of the Middle Cambrian Phyllopod Bed (Burgess Shale): Palaeontology 29(3):423–467. Google Scholar

    61.

    Conway Morris, S. 1992. Burgess Shale-type faunas in the context of the ‘Cambrian explosion: a review: Journal of the Geological Society 149(4):631–636. Google Scholar

    62.

    Conway Morris, S., & R. A. Robison. 1986. Middle Cambrian priapulids and other soft-bodied fossils from Utah and Spain: University of Kansas Paleontological Contributions 117:1–22. Google Scholar

    63.

    Conway Morris, S., & R. A. Robison. 1988. More soft-bodied animals and algae from the Middle Cambrian of Utah and British Columbia: The University of Kansas Paleontological Contributions 122(1):1–48. Google Scholar

    64.

    Counts, J. W., & S. T. Hasiotis. 2009. Neoichnological experiments with masked chafer beetles (Coleoptera: Scarabaeidae): Implications for backfilled continental trace fossils: Palaios 24(2):74–91. Google Scholar

    65.

    Crimes, T. P. 1970a. Trilobite tracks and other trace fossils from the Upper Cambrian of North Wales: Geological Journal 7(1):47–68. Google Scholar

    66.

    Crimes, T. P. 1970b. The significance of trace fossils in sedimentology, stratigraphy, and palaeontology with examples from Lower Palaeozoic strata. In T. P. Crimes, & J. C. Harper, eds., Trace fossils. Geological Journal Special Publication 3. p. 101–126. Google Scholar

    67.

    Crimes, T. P. 1975. Trilobite traces from the Lower Tremadoc of Tortworth: Geological Magazine 112(1):33–46. Google Scholar

    68.

    Crimes, T. P. 1987. Trace fossils and correlation of late Precambrian and early Cambrian strata: Geological Magazine 124(2):97–119. Google Scholar

    69.

    Crimes, T. P. 1992. Changes in the trace fossil biota across the Proterozoic-Phanerozoic boundary: Journal of the Geological Society 149(4):637–646. Google Scholar

    70.

    Crimes, T. P., & M. M. Anderson. 1985. Trace Fossils from Late Precambrian-Early Cambrian Strata of Southeastern Newfoundland (Canada): Temporal and Environmental Implications: Journal of Paleontology 59(2):310–343. Google Scholar

    71.

    Crimes, T. P., & M. A. Fedonkin. 1994. Evolution and Dispersal of Deepsea Traces: Palaios 9(1):74–83. Google Scholar

    72.

    Crimes, T. P., & G. J. B. Germs. 1982. Trace fossils from the Nama Group (Precambrian—Cambrian) of southwest Africa (Namibia): Journal of Paleontology 56(4):890–907. Google Scholar

    73.

    Crimes, T. P., R. Goldring, P. Homewood, J. Van Stuijvenberg, & W. Winkler. 1981. Trace fossil assemblages of deep-sea fan deposits, Gurnigel and Schlieren flysch (Cretaceous-Eocene), Switzerland: Eclogae Geologicae Helvetiae 74(3):953–995. Google Scholar

    74.

    Crimes, T. P., I. Legg, A. Marcos, & M. Arboleya. 1977. ?Late Precambrian— low Lower Cambrian trace fossils from Spain. In T. P. Crimes, & J. C. Harper, eds., Trace Fossils 2: Proceedings from the 25th International Geological Conference. Geological Journal Special Issue 9. p. 91–138. Google Scholar

    75.

    Curran, H. A., & R. W. Frey. 1977. Pleistocene trace fossils from North Carolina (USA), and their Holocene analogues. In T. P. Crimes, & J. C. Harper, eds., Trace Fossils 2: Proceedings from the 25th International Geological Conference. Geological Journal Special Issue 9. p. 139–162. Google Scholar

    76.

    Dahmer, G. 1937. Lebensspuren aus dem Taunusquarzit und den Siegener Schichten (Unterdevon): Preussische Geologische Ladesanstalt, Jahrbuch , 1936, 57:523–539, text-fig. 1–2, pl. 31–35. Google Scholar

    77.

    D'Alessandro, A., & R. G. Bromley. 1987. Meniscate trace fossils and the Muensteria-Taenidium problem: Palaeontology 30(4):743–763. Google Scholar

    78.

    Dawson, J. W. 1873. Impressions and footprints of aquatic animals and imitative markings on Carboniferous rocks: American Journal of Science (25):16–24. Google Scholar

    79.

    Dawson, S. J. W. 1864. On the fossils of the genus Rusophycus: The Canadian Naturalist and Geologist 2:363–367. Google Scholar

    80.

    DeFelice, R. C., L. G. Eldredge, & J. T. Carlton. 2001. Non-indigenous Invertebrates. In L. G. Eldredge, and C. M. Smith, eds., A Guidebook of Introduced Marine Species in Hawai'i. Hawai'i Biological Survey. Honolulu. Bishop Museum Technical Report 21. p. B-1—B-60. Google Scholar

    81.

    De Gibert, J. M., M. A. Fregenal-Martínez, L. A. Buatois, & M. G. Mángano. 2000. Trace fossils and their palaeoecological significance in Lower Cretaceous lacustrine conservation deposits, El Montsec, Spain: Palaeogeography, Palaeoclimatology, Palaeoecology 156(1):89–101. Google Scholar

    82.

    De Stefani, C. 1885. Studi paleozoologici sulle creta superiore e media dell'Appennino settentrionale: Memorie della Classe di Scienze Fisiche, Matemariche e Naturali 4(1):73–121. Google Scholar

    83.

    Devera, J. A. 1989. Ichnofossil assemblages and associated lithofacies of the Lower Pennsylvanian (Caseyville and Tradewater Formations), southern Illinois: Illinois Basin Studies 1:57–83. Google Scholar

    84.

    d'Orbigny, A. D. 1842. Voyage dans l'Amerique méridionale (le Bresil, la Repúblique argentine, le Patagonia, Le Repúblique de Chile, le Repúblique de Bolivie, Le Repúblique de Perou) executé pendant les années 1826–1833. P. Bertrand. Paris. 3 (Paleontologie). 188 p. Google Scholar

    85.

    Droser, M. L., & D. J. Bottjer. 1986. A semiquantitative field classification of ichnofabric: Research method paper: Journal of Sedimentary Research 56(4):558–559. Google Scholar

    86.

    Droser, M. L., S. Jensen, J. G. Gehling, P. M. Myrow, & G. M. Narbonne. 2002. Lowermost Cambrian Ichnofabrics from the Chapel Island Formation, Newfoundland: Implications for Cambrian Substrates, Palaios 17:3–15. Google Scholar

    87.

    Dzik, J. 2005. Behavioral and anatomical unity of the earliest burrowing animals and the cause of the “Cambrian explosion”: Paleobiology 31(3):503–521. Google Scholar

    88.

    Eagar, R. M. C. 1974. Shape of shell of Carbonicola in relation to burrowing: Lethaia 7(3):219–238. Google Scholar

    89.

    Ehrenburg, K. 1944. Ergänzende Bemerkungen zu den seiner zest aus dem Miozän von Burgschlienitz beschriebenen Gangkernen und Bauten dekapoder Krebse: Paläontologische Zeitschrift 23:354–359. Google Scholar

    90.

    Eichwald, C. E. v. 1860. Lethaea rossica ou Paléontologie de la Russie. E. Schweizerbart. Stuttgart. 1657 p. Google Scholar

    91.

    Ekdale, A. A., R. G. Bromley, & D. B. Loope. 2007. Ichnofacies of an ancient erg: A climatically influenced trace fossil association in the Jurassic Navajo Sandstone, southern Utah, USA. In W. Miller III, ed., Trace fossils: concepts, problems, prospects. Elsevier. Amsterdam, p. 562–574. Google Scholar

    92.

    Ekdale, A. A., R. G. Bromley, & S. G. Pemberton. 1984. Ichnology: The Use of Trace Fossils in Sedimentology and Stratigraphy. SEPM Short Course 15. Tulsa, Oklahoma. 317 p. Google Scholar

    93.

    Ekdale, A. A., & T. R. Mason. 1988. Characteristic trace-fossil associations in oxygen-poor sedimentary environments: Geology 16:720–723. Google Scholar

    94.

    Elliott, D. K., & D. L. Martin. 1987. A new trace fossil from the Cambrian Bright Angel Shale, Grand Canyon, Arizona: Journal of Paleontology 61(4):641–648. Google Scholar

    95.

    Emmons, E. 1844. The Taconic System: Based on Observations in New York, Massachusetts, Maine, Vermont and Rhode Island. Carroll and Cook Printers. Albany, New York. 68 p. Google Scholar

    96.

    Feibel, C. S. 1987. Fossil fish nests from the Koobi Fora Formation (Plio-Pleistocene) of northern Kenya: Journal of Paleontology 61(1):130–134. Google Scholar

    97.

    Fenton, C. L., & M. A. Fenton. 1937a. Archaeonassa: Cambrian snail trails and burrows: American Midland Naturalist 18(3):454–456. Google Scholar

    98.

    Fenton, C. L., & M. A. Fenton. 1937b. Burrows and trails from Pennsylvanian rocks of Texas: American Midland Naturalist 18(6):1079–1084. Google Scholar

    99.

    Fenton, C. L., & M. A. Fenton. 1937c. Olivellites, a Pennsylvanian Snail Burrow: American Midland Naturalist 18(3):452–453. Google Scholar

    100.

    Fenton, C. L., & M. A. Fenton. 1937d. Trilobite “Nests” and Feeding Burrows: American Midland Naturalist 18(3):446–451. Google Scholar

    101.

    Fernández, D. E., P. J. Pazos, & M. B. Aguirre-Urreta. 2010. Protovirgularia dichotoma—Protovirgularia rugosa: An Example of a Compound Trace Fossil from the Lower Cretaceous (Agrio Formation) of the Neuquén Basin, Argentina: Ichnos 17:40–47. Google Scholar

    102.

    Fillion, D., & R. K. Pickerill. 1984. Systematic ichnology of the Middle Ordovician Trenton Group, St Lawrence Lowland, eastern Canada: Atlantic Geology 20(1):1–41. Google Scholar

    103.

    Fillion, D., & R. K. Pickerill. 1990. Ichnology of the Upper Cambrian? to Lower Ordovician Bell Island and Wabana groups of eastern Newfoundland, Canada. Palaeontographica Canadiana 7. Canadian Society of Petroleum Geologists, Geological Association of Canada. 119 p. Google Scholar

    104.

    Fischer-Ooster, C. v. 1858. Die fossilen Fucoiden der Schweizer-Alpen, nebst Erörterungen über deren geologisches Alter. Huber und Companie. Bern. 74 p. Google Scholar

    105.

    Fisher, W. A. 1978. Trace fossils from the Harding Formation (Middle Ordovician), Colorado, Rocky Mountain Association Geologists 1978 Field Conference and Symposium of Energy Resources of the Denver Basin Guidebook. Rocky Mountain Association Geologists. Denver, p. 191–197. Google Scholar

    106.

    Fitch, A. 1850. A historical, topographical, and agricultural survey of the County of Washington Transactions of the New York Agricultural Society 9:753–944. Google Scholar

    107.

    Forster, S. 1996. Spatial and Temporal Distribution of Oxidation Events Occurring Below the Sediment—Water Interface: Marine Ecology 17 (1–3):309–319. Google Scholar

    108.

    Fortey, R. A., & A. Schacher. 1997. The trace fossil Cruziana semiplicata and the trilobite that made it: Lethaia 30(2):105–112. Google Scholar

    109.

    Frey, R. W., & J. D. Howard. 1981. Conichnus and Schaubcylindrichnus: Redefined Trace Fossils from the Upper Cretaceous of the Western Interior: Journal of Paleontology 55(4):800–804. Google Scholar

    110.

    Frey, R. W., & J. D. Howard. 1985. Trace fossils from the Panther Member, Star Point Formation (Upper Cretaceous), Coal Creek Canyon, Utah: Journal of Paleontology 59:370–404. Google Scholar

    111.

    Fu, S. 1991. Funktion, Verhalten und Einteilung fucoider und lophoctenoider Lebensspuren: Courier Forschungs-Institut Senckenberg 135:1–79. Google Scholar

    112.

    Fu, S., & F. Werner. 2000. Distribution, ecology and taphonomy of the organism trace, Scolicia, in northeast Atlantic deep-sea sediments: Palaeogeography, Palaeoclimatology, Palaeoecology 156(3–4): 289–300. Google Scholar

    113.

    Fürsich, F. T. 1974a. On Diplocraterion Torell 1870 and the Significance of Morphological Features in Vertical, Spreiten-Bearing, U-Shaped Trace Fossils: Journal of Paleontology 48(5):952–962. Google Scholar

    114.

    Fürsich, F. T. 1974b. Corallian (Upper Jurassic) trace fossils from England and Normandy: Stuttgarter Beiträge zur Naturkunde (Serie B) Geologie und Paläontologie 13:1–51. Google Scholar

    115.

    Fürsich, F. T. 1975. Trace fossils as environmental indicators in the Corallian of England and Normandy: Lethaia 8(2):151–172. Google Scholar

    116.

    Fürsich, F. T., & R. G. Bromley. 1985. Behavioural interpretation of a rosetted spreite trace fossil: Dactyloidites otto (Geinitz): Lethaia 18(3):199–207. Google Scholar

    117.

    Fürsich, F. T., & W. J. Kennedy. 1975. Kirklandia texana Caster—Cretaceous hydrozoan medusoid or trace fossil chimaera?: Palaeontology 18(4):665–679. Google Scholar

    118.

    Gabelli, L. d. 1900. Sopra un interessante impronta medusoide: Il Pensiero Aristotelico nella Scienza Moderna 1(2):74–79. Google Scholar

    119.

    Gaillard, C., J. Goy, P. Bernier, J. P. Bourseau, J. C. Gall, G. Barale, E. Buffetaut, & S. Wenz. 2006. New jellyfish taxa from the upper Jurassic lithographic limestones of Germ (France): Taphonomy and ecology: Palaeontology 49(6):1287–1302. Google Scholar

    120.

    Gaillard, C., P. Hantzpergue, J. Vannier, A.-L. Margerard, & J.-M. Mazin. 2005. Isopod Trackways from the Crayssac Lagerstätte, Upper Jurassic, France: Palaeontology 48(5):947–962. Google Scholar

    121.

    Gaillard, C., & D. Olivero. 2009. The ichnofossil Halimedides in Cretaceous pelagic deposits from the Alps: Environmental and ethological significance Palaios 24(4):257–270. Google Scholar

    122.

    Gaines, R. R., & M. L. Droser. 2005. New Approaches to Understanding the Mechanics of Burgess Shale-type Deposits: From the Micron Scale to the Global Picture: The Sedimentary Record 3(2):4–8. Google Scholar

    123.

    Gaines, R. R., & M. L. Droser. 2010. The paleredox setting of Burgess Shale-type deposits: Palaeogeography, Palaeoclimatology, Palaeoecology 297(3–4):649–661. Google Scholar

    124.

    Gaines, R. R., M. J. Kennedy, & M. L. Droser. 2005. A new hypothesis for organic preservation of Burgess Shale taxa in the middle Cambrian Wheeler Formation, House Range, Utah: Palaeogeography, Palaeoclimatology, Palaeoecology 220(1–2):193–205. Google Scholar

    125.

    Garson, D. E., R. R. Gaines, M. L. Droser, W. D. Liddell, & A. Sappenfield. 2012. Dynamic palaeoredox and exceptional preservation in the Cambrian Spence Shale of Utah: Lethaia 45(2):164–177. Google Scholar

    126.

    Garvey, J. M., & S. T. Hasiotis. 2008. An ichnofossil assemblage from the Lower Carboniferous Snowy Plains Formation, Mansfield Basin, Australia: Palaeogeography, Palaeoclimatology, Palaeoecology 258(4):257–276. Google Scholar

    127.

    Gehling, J. G., S. Jensen, M. L. Droser, P. M. Myrow, & G. M. Narbonne. 2001. Burrowing below the basal Cambrian GSSP, Fortune Head, Newfoundland: Geological Magazine 138(2):213–218. Google Scholar

    128.

    Germs, G. J. B. 1972. Trace Fossils from the Nama Group, South-West Africa: Journal of Paleontology 46(6):864–870. Google Scholar

    129.

    Getty, P. R., T. D. McCarthy, S. Hsieh, & A. M. Bush. 2016. A new reconstruction of continental Treptichnus based on exceptionally preserved material from the Jurassic of Massachusetts: Journal of Paleontology 90(2):269–278. Google Scholar

    130.

    Gevers, T. W. 1973. A New Name for the Ichnogenus Arthropodichnus Gevers, 1971: Journal of Paleontology 47(5):1002. Google Scholar

    131.

    Gevers, T. W., L. A. Frakes, L. N. Edwards, & J. E. Marzolf. 1971. Trace fossils in the lower Beacon sediments (Devonian), Darwin mountains, southern Victoria Land, Antarctica: Journal of Paleontology 45(1):81–94. Google Scholar

    132.

    Geyer, G., & A. Uchman. 1995. Ichnofossil assemblages from the Nama Group (Neoproterozoic-Lower Cambrian) in Namibia and the Proterozoic-Cambrian boundary problem revisited. In G. Geyer, & E. Landing, eds., Morocco ‘95: The Lower—Middle Cambrian Standard of Western Gondwana. Beringeria Special Issue 2. p. 175–202. Google Scholar

    133.

    Gibb, S., B. D. E. Chatterton, & S. G. Pemberton. 2009. Arthropod ichnofossils from the Ordovician Stairway Sandstone of central Australia: Association of Australasian Palaeontologists Memoirs 37:695–716. Google Scholar

    134.

    Gingras, M. K., K. L. Bann, J. A. MacEachern, & S. G. Pemberton. 2007. A Conceptual Framework for the Application of Trace Fossils. In J. A. MacEachern, K. L. Bann, M. K. Gingras, and S. G. Pemberton, eds., Applied Ichnology: SEPM Short Course Notes 52. SEPM 52:1–26. Google Scholar

    135.

    Gingras, M. K., J. A. MacEachern, S. E. Dashtgard, M. J. Ranger, & S. G. Pemberton. 2016. The signifcance of trace fossils in the McMurray Formation, Alberta, Canada: Bulletin of Canadian Petroleum Geology 64(2):233–250. Google Scholar

    136.

    Gingras, M. K., J. A. MacEachern, & S. G. Pemberton. 1998. A comparative analysis of the ichnology of wave-and river-dominated allomembers of the Upper Cretaceous Dunvegan Formation: Bulletin of Canadian Petroleum Geology 46(1):51–73. Google Scholar

    137.

    Glocker, F. E. 1841. Über die kalkführende Sandsteinformation auf beiden Seiten der mittleren March, in der Gegend zwischen Kwassitz und Kremsier: Academia Caesarea Leopoldino-Carolina Germánica Naturae Curiosorum 19(Supplement 2):309–334. Google Scholar

    138.

    Głuszek, A. 1995. Invertebrate trace fossils in the continental deposits of an Upper Carboniferous coal-bearing succession, Upper Silesia, Poland: Studia Geologica Polonica 108:171–202. Google Scholar

    139.

    Głuszek, A. 1998. Trace fossils from Late Carboniferous storm deposits, Upper Silesia Coal Basin, Poland: Acta Palaeontologica Polonica 43(3):517–546. Google Scholar

    140.

    Goldring, R. 1962. The trace fossils of the Baggy beds (upper Devonian) of north Devon, England: Paläontologische Zeitschrift 36(3–4) :232–251. Google Scholar

    141.

    Götzinger, G., & H. Becker. 1932. Zur geologischen Gliederung des Wienerwaldflysches (neue Fossilfunde) Geologischen Bundesanstalt Wien Jahrbuch 82:343–396. Google Scholar

    142.

    Gunther, L. F., & V. G. Gunther. 1981. Some Middle Cambrian fossils of Utah: Brigham Young University Geology Studies 28(1):1–81. Google Scholar

    143.

    Guthörl, P. 1934. Die Arthropoden aus dem Carbon und Perm des Saar-Nahe-Pjalz-Gebietes: Preussische Geologische Landesantalt, Abhandlungen , new series, 164:219 p., text-fig. 1–116, 30 pl. Google Scholar

    144.

    Hagadorn, J. W. 2002. Burgess Shale: Cambrian explosion in full bloom. In D. J. Bottjer, W. Etter, J. W. Hagadorn, & T. C. M., eds., Exceptional Fossil Preservation: A Unique View on the Evolution of Marine Life. Columbia University Press. New York. p. 61–89. Google Scholar

    145.

    Hagadorn, J. W., R. H. Dott, & D. Damrov. 2002. Stranded on a Late Cambrian shoreline: Medusae from central Wisconsin: Geology 30(2):147–150. Google Scholar

    146.

    Hagdorn, H. 2014. 12. Spurenfossilien aus dem Lettenkeuper. In H. Hagdorn, R. Schoch, G. Schweigert, eds., Der Lettenkeuper - ein Fenster in die Zeit vor den Dinosauriern. Palaeodiversity Special Issue, p. 267–281. Google Scholar

    147.

    Hakes, W. G. 1976. Trace fossils and depositional environment of four clastic units, Upper Pennsylvanian megacyclothems, northeast Kansas: University of Kansas Paleontological Contributions 63:1–46, pl. 13. Google Scholar

    148.

    Hakes, W. G. 1977. Trace fossils in Late Pennsylvanian cyclothems, Kansas. In T. P. Crimes, and J. C. Harper, eds., Trace Fossils 2: Proceedings from the 25th International Geological Conference. Geological Journal. Special Issue 9. p. 209–226. Google Scholar

    149.

    Hakes, W. G. 1985. Trace fossils from brackish-marine shales, Upper Pennsylvanian of Kansas, USA. In H. A. Curran, ed., Biogenic Structures: Their Use in Interpreting Depositional Environments. The Society of Economic Paleontologists and Mineralogists. Special Publication 35:21–35. Google Scholar

    150.

    Hall, J. 1847. Palaeontology of New York. Volume I. Containing descriptions of the organic remains of the Lower Division of of the New York System (equivalent of the Lower Silurian rocks of Europe). C. van Benthuysen. Albany, New York. 338 p. Google Scholar

    151.

    Hall, J. 1852. Palaeontology of New York. Volume II. Containing descriptions of the organic remains of the Lower Middle Division of the New York System, (Equivalent in part to the Middle Silurian rocks of Europe). C. van Benthuysen. Albany, New York. 362 p. Google Scholar

    152.

    Hall, J. 1886. Note on some obscure organisms in the roofing slate of Washington County, New York: New York State Museum of Natural History, Annual Report 39:160. Google Scholar

    153.

    Hallam, A. 1970. Gyrochorte and other trace fossils in the Forest Marble (Bathonian) of Dorset, England. In T. P. Crimes, & J. C. Harper, eds., Trace fossils. Geological Journal Special Publication 3:189–200. Google Scholar

    154.

    Han, Y., & R. K. Pickerill. 1994a. Phycodes templus isp. nov. from the Lower Devonian of northwestern New Brunswick, eastern Canada: Atlantic Geology 30:37–46. Google Scholar

    155.

    Han, Y., & R. K. Pickerill. 1994b. Taxonomic reassessment of Protovirguiaria M'Coy 1850 with new examples from the Paleozoic of New Brunswick, eastern Canada: Ichnos 3(3):203–212. Google Scholar

    156.

    Häntzschel, W. 1975. Trace Fossils and Problematica. In C. Teichert, ed., Treatise on Invertebrate Paleontology. Part W: Miscellanea (Supplement 1). University of Kansas Press & Geological Society of America. Lawrence, Kansas & Boulder, Colorado. 269 p. Google Scholar

    157.

    Harris, B. S., E. R. Timmer, M. J. Ranger, & M. K. Gingras. 2016. Continental ichnology of the Lower McMurray Formation inclined heterolithic strata at Daphne Island, Athabasca River, northeastern Alberta, Canada: Bulletin of Canadian Petroleum Geology 64(2):218–232. Google Scholar

    158.

    Hasioris, S. T. 2002. Continental Trace Fossils. SEPM Short Course 51. Tulsa, Oklahoma. 132 p. Google Scholar

    159.

    Hasioris, S. T. 2004. Reconnaissance of Upper Jurassic Morrison Formation ichnofossils, Rocky Mountain Region, USA: paleoenvironmental, stratigraphic, and paleoclimatic significance of terrestrial and freshwater ichnocoenoses: Sedimentary Geology 167(3):177–268. Google Scholar

    160.

    Hasiotis, S. T. 2007. Continental ichnology: fundamental processes and controls on trace-fossil distribution. In W. Miller, III, ed., Trace Fossils—concepts, problems, prospects. Elsevier. Amsterdam, p. 268–284. Google Scholar

    161.

    Hasiotis, S. T. 2008. Reply to the Comments by Bromley et al. of the paper “Reconnaissance of the Upper Jurassic Morrison Formation ichnofossils, Rocky Mountain Region, USA: Paleoenvironmental, stratigraphic, and paleoclimatic significance of terrestrial and freshwater ichnocoenoses” by Stephen T. Hasiotis: Sedimentary Geology 208(1):61–68. Google Scholar

    162.

    Hasioris, S. T. 2012. A Brief Overview of the Diversity and Patterns in Bioturbation Preserved in the Cambrian—Ordovician Carbonate and Siliciclastic Deposits of Laurentia. In J. R. Derby, R. D. Fritz, S. A. Longacre, W. A. Morgan, & C. A. Sternbach, eds., The great American carbonate bank: The geology and economic resources of the Cambrian—Ordovician Sauk megasequence of Laurentia. AAPG Memoir 98. p. 111–123. Google Scholar

    163.

    Hasioris, S. T., & C. E. Mitchell. 1993. A comparison of crayfish burrow morphologies: Triassic and Holocene fossil, paleo- and neo-ichnological evidence , and the identification of their burrowing signatures: Ichnos 2(4):291–314. Google Scholar

    164.

    Hasioris, S. T., C. E. Mitchell, & R. F. Dubiel. 1993. Application of morphologic burrow interpretations to discern continental burrow architects: Lungfish or crayfish?: Ichnos 2(4):315–333. Google Scholar

    165.

    Hasioris, S. T., & B. F. Platt. 2012. Exploring the sedimentary, pedogenic, and hydrologic factors that control the occurrence and role of bioturbation in soil formation and horizonation in continental deposits: An integrative approach: The Sedimentary Record 10(3):4–9. Google Scholar

    166.

    Hasiotis, S. T., B. F. Platt, M. Reilly, K. Amos, S. Lang, D. Kennedy, J. A. Todd, & E. Michel. 2012. Actualistic studies of the spatial and temporal distribution of terrestrial and aquatic organism traces in continental environments to differentiate lacustrine from fluvial, eolian, and marine deposits in the geologic record. In O. W. Baganz, Y. Bartov, K. M. Bohacs, & D. Nummedal, eds., Lacustrine sandstone reservoirs and hydrocarbon systems. AAPG Memoir 95. p. 433–489. Google Scholar

    167.

    Heer, O. 1864–1865. Die urwelt der Schweiz. F. Schulthess. Zürich. 622 p. Google Scholar

    168.

    Heer, O. 1877. Flora fossilis Helvetiae: Die vorweltliche flora der Schweiz. J. Wurster & Company. Zürich. 182 p. Google Scholar

    169.

    Heinberg, C. 1973. The internal structure of the trace fossils Gyrochorte and Curvolithus: Lethaia 6:227–238. Google Scholar

    170.

    Heinberg, C. 1974. A dynamic model for a meniscus filled tunnel (Ancorichnus n. ichnogen.) from the Jurassic Pecten Sandstone of Milne Land, East Greenland: Rapport Grønlands Geologiske Undersøgelse 62:1–20. Google Scholar

    171.

    Hembree, D. I., & S. T. Hasiotis. 2007. Paleosols and ichnofossils of the White River Formation of Colorado: Insight into soil ecosystems of the North American Midcontinent during the Eocene-Oligocene transition: Palaios 22(2):123–142. Google Scholar

    172.

    Hintze, L. F., & R. A. Robison. 1975. Middle Cambrian Stratigraphy of the House, Wah Wah, and Adjacent Ranges in Western Utah: Geological Society of America Bulletin 86(7):881–891. Google Scholar

    173.

    Hiscott, R. N., N. P. James, & S. G. Pemberton. 1984. Sedimentology and ichnology of the Lower Cambrian Bradore Formation, coastal Labrador: fluvial to shallow-marine transgressive sequence: Bulletin of Canadian Petroleum Geology 32(1):11–26. Google Scholar

    174.

    Hitchcock, E. 1858. Ichnology of New England: a Report on the Sandstone of the Connecticut Valley Especially Its Fossil Footmarks Made to the Government of the Commonwealth of Massachusetts. White. Boston. 220 p. Google Scholar

    175.

    Hitchcock, E. 1865. Supplement to the Ichnology of New England: A Report to the Government of Massachusetts, in 1863. Wright & Potter. Boston. 96 p. Google Scholar

    176.

    Hofmann, R., M. G. Mángano, O. Elicki, & R. Shinaq. 2012. Paleoecologic and biostratigraphic significance of trace fossils from shallow-to marginal-marine environments from the middle Cambrian (Stage 5) of Jordan: Journal of Paleontology 86(6):931–955. Google Scholar

    177.

    Hu, B., G. Wang, & R. Goldring. 1998. Nereites (or Neonereites) from lower Jurassic lacustrine turbidites of Henan, central China: Ichnos 6(3):203–209. Google Scholar

    178.

    Huang, D., J. Chen, M. Zhu, & F. Zhao. 2014. The burrow dwelling behavior and locomotion of palaeoscolecidian worms: New fossil evidence from the Cambrian Chengjiang fauna: Palaeogeography, Palaeoclimatology, Palaeoecology 398:154–164. Google Scholar

    179.

    Hummelinck, P. W. 1968. Caribbean scyphomedusae of the genus Cassiopea: Studies on the Fauna of Curacao and other Caribbean Islands 25(1):1–57. Google Scholar

    180.

    ICZN (International Code of Zoological Nomenclature). 1964. International Code of Zoological Nomenclature, 2nd edition. The International Trust for Zoological Nomenclature. London. 176 p. Google Scholar

    181.

    Ineson, J. R., & J. S. Peel. 2011. Geological and depositional setting of the Sirius Passet Lagerstätte (Early Cambrian), North Greenland: Canadian Journal of Earth Sciences 48(8):1259–1281. Google Scholar

    182.

    Jackson, A. M., S. T. Hasiotis, & P. P. Flaig. 2016. Ichnology of a Paleopolar, River-Dominated, Shallow Marine Deltaic Succession ín the Mackellar Sea: The Mackellar Formation (Lower Permian), Central Transantarctic Mountains, Antarctica: Palaeogeography, Palaeoclimatology, Palaeoecology 441(2):266–291. Google Scholar

    183.

    James, U. P. 1879. Description of new species of fossils and remarks on some others, from the Lower and Upper Silurian rocks of Ohio: Palaeontologist 3:17–24. Google Scholar

    184.

    Jenkins, R. J. F. 1995. The problems and potential of using animal fossils and trace fossils in terminal Proterozoic biostratigraphy: Precambrian Research 73:51–69. Google Scholar

    185.

    Jensen, M. E., & J. K. King. 1999. Geologic map of the Brigham City 7.5-minute quadrangle, Box Elder and Cache Counties, Utah. Utah Geological Survey. Salt Lake City. Map 173. Google Scholar

    186.

    Jensen, S. 1990. Predation by early Cambrian trilobites on infaunal worms-evidence from the Swedish Mickwitzia Sandstone: Lethaia 23(1):29–42. Google Scholar

    187.

    Jensen, S. 1997. Trace fossils from the Lower Cambrian Mickwitzia sandstone, south-central Sweden. Fossils and Strata 42. Scandinavian University Press. Oslo. 111 p. Google Scholar

    188.

    Jensen, S., M. L. Droser, & J. G. Gehling. 2005. Trace fossil preservation and the early evolution of animals: Palaeogeography, Palaeoclimatology, Palaeoecology 220(1–2):19–29. Google Scholar

    189.

    Jensen, S., M. L. Droser, & J. G. Gehling. 2006. A critical look at the Ediacaran trace fossil record. In S. Xiao & A. J. Kaufman, eds., Neop rote rozoic Geobiology and Paleobiology. Springer Netherlands. Amsterdam, p. 115–157. Google Scholar

    190.

    Jensen, S., M. L. Droser, & N. A. Heim. 2002. Trace fossils and ichnofabrics of the Lower Cambrian Wood Canyon Formation, southwest Death Valley area. In F. A. Corsetti, ed., Proterozoic—Cambrian of the Great Basin and Beyond: Volume and Guidebook for the Society for Sedimentary Geology (SEPM) 93:123–135. Google Scholar

    191.

    Jensen, S., B. Z. Saylor, J. G. Gehling, & G. J. B. Germs. 2000. Complex trace fossils from the terminal Proterozoic of Namibia: Geology 28(2): 143–146. Google Scholar

    192.

    Jones, M. F., 2016. Neoichnology of bats: Morphological, Ecological, and Phylogenetic Influences on Terrestrial Behavior and Trackmaking Ability within the Chiroptera. Unpublished M.S. thesis. University of Kansas. Lawrence, Kansas. 159 p. Google Scholar

    193.

    Karaszewski, W. 1975. A new trace fossil from the Lower Jurassic of the Holy Cross Mountains: Bulletin de l'Académie Polonaise des Sciences. Série des Sciences de la Terre 22:157–160. Google Scholar

    194.

    Kawase, H., Y. Okata, & K. Ito. 2013. Role of Huge Geometric Circular Structures in the Reproduction of a Marine Pufferfish: Scientific Reports 3(2106): 1–5. Google Scholar

    195.

    Keighley, D. G., & R. K. Pickerill. 1994. The ichnogenus Beaconites and its distinction from Ancorichnus and Taenidium: Palaeontology 37(2):305–338. Google Scholar

    196.

    Keighley, D. G., & R. K. Pickerill. 1995. Commentary: The ichnotaxa Palaeophycus and Planolites: Historical perspectives and recommendations: Ichnos 3(4):301–309. Google Scholar

    197.

    Keighley, D. G., & R. K. Pickerill. 1996. Small Cruziana, Rusophycus, and related ichnotaxa from eastern Canada: the nomenclatural debate and systematic ichnology: Ichnos, 4(4): 261–285. Google Scholar

    198.

    Keighley, D. G., & R. K. Pickerill. 1997. Systematic ichnology of the Mabou and Cumberland groups (Carboniferous) of western Cape Breton Island, eastern Canada, 1: burrows, pits, trails, and coprolites: Atlantic Geology 33(3): 181–215. Google Scholar

    199.

    Keighley, D. G., & R. K. Pickerill. 1998. Systematic ichnology of the Mabou and Cumberland groups (Carboniferous) of western Cape Breton Island, eastern Canada, 2: surface markings: Atlantic Geology 34(2):83–112. Google Scholar

    200.

    Keighley, D. G., & R. K. Pickerill. 2003. Ichnocoenoses from the Carboniferous of eastern Canada and their implications for the recognition of ichnofacies in nonmarine strata: Atlantic Geology 39(1): 1–22. Google Scholar

    201.

    Keller, B. M., V. V. Menner, V. A. Stepanov, & N. M. Chumakov. 1974. [New finds of fossils in the Precambrian Valdai Series along the Syuzma River] (in Russian): Izvestia Akademii Nauk SSSR, Seriya Geologiya 12:130–134. Google Scholar

    202.

    Kern, J. P. 1978. Paleoenvironment of new trace fossils from the Eocene Mission Valley Formation, California: Journal of Paleontology 52:186–194. Google Scholar

    203.

    Khaidem, K. S., H. S. Rajkumar, & I. Soibam. 2015. Attribute of trace fossils of Laisong flysch sediments, Manipur, India: Journal of Earth System Science 124(5): 1085–1113. Google Scholar

    204.

    Kim, J.-Y. 1994. A unique occurrence of Lockeia from the Yeongheung Formation (Middle Ordovician), Yeongweol, Korea: Ichnos 3:219–225. Google Scholar

    205.

    Kim, J. Y., D. G. Keighley, R. K. Pickerill, W. Hwang, & K.-S. Kim. 2005. Trace fossils from marginal lacustrine deposits of the Cretaceous Jinju Formation, southern coast of Korea: Palaeogeography, Palaeocfimatology, Palaeoecology 218(1–2): 105–124. Google Scholar

    206.

    Kim, K.-S., & J.-Y. Kim. 2008. Lockeia gigantas ichnosp. nov. in the Lacustrine Deposits of the Early Cretaceous Jinju Formation, Southern Coast of Korea: Journal of the Korean Earth Science Society 29(1): 13–28. Google Scholar

    207.

    Knaust, D. 2007. Invertebrate trace fossils and ichnodiversity in shallowmarine carbonates of the German Middle Triassic (Muschelkalk). In R. Bromley, L. A. Buatois, G. Mángano, J. F. Genise, & R. N. Melchor, eds., Sediment-Organism Interactions: A Multifaceted Ichnology. Society for Sedimentary Geology. Tulsa, Oklahoma. Special Publication 88. p. 223–240. Google Scholar

    208.

    Knaust, D., M. J. Warchoł, & I. A. Kane. 2014. Ichnodiversity and ichnoabundance: Revealing depositional trends in a confined turbidite system: Sedimentology 61:2218–2267. Google Scholar

    209.

    Kramer, J. M., B. R. Erickson, M. G. Lockley, A. P. Hunt, & S. J. Braddy. 1995. Pelycosaur predation in the Permian: Evidence from Laoporus trackways from the Coconino Sandstone with description of a new species of Permichnium. In S. G. Lucas, & A. B. Heckert, eds., Early Permian foot prints and facies. New Mexico Museum of Natural History and Science Bulletin 6. p. 245–249. Google Scholar

    210.

    Krapovickas, V., P. L. Ciccioli, M. G. Mángano, C. A. Marsicano, & C. O. Limarino. 2009. Paleobiology and paleoecology of an arid—semiarid Miocene South American ichnofauna in anastomosed fluvial deposits: Palaeogeography, Palaeoclimatology, Palaeoecology 284(3):129–152. Google Scholar

    211.

    Książkiewicz, M. 1977. Trace fossils in the flysch of the Polish Carpathians: Palaeontologia Polonica 36:1–270. Google Scholar

    212.

    Kukal, Z. 1995. The Lower Cambrian Paseky Shale: Sedimentology: Journal of the Czech Geological Society 40(4):67–78. Google Scholar

    213.

    Landing, E., S. Peng, L. E. Babcock, G. Geyer, & M. MoczydlowskaVidal. 2007. Global standard names for the Lowermost Cambrian Series and Stage: Episodes 30(4):287–289. Google Scholar

    214.

    Lane, A. A., S. J. Braddy, D. E. G. Briggs, & D. K. Elliott. 2003. A new trace fossil from the Middle Cambrian of the Grand Canyon, Arizona, USA: Palaeontology 46(5):987–997. Google Scholar

    215.

    Lapworth, C. 1870. On the Lower Silurian rocks in the neighbourhood of Galashiels: Transactions of the Royal Society of Edinburgh 2:46–58. Google Scholar

    216.

    Le Roux, J. P., S. N. Nielson, & A. Henriquez. 2008. Depositional environment of Stelloglyphus llicoensis isp. nov.: a new radial trace fossil from the Neogene Ranquil Formation, south-central Chile: Revista Geológica de Chile 35(2):307–319. Google Scholar

    217.

    Legg, I. C. 1985. Trace fossils from a Middle Cambrian deltaic sequence, north Spain. In H. A. Curran, ed., Biogenic Structures: Their Use in Interpreting Depositional Environments. SEPM. Tulsa, Oklahoma. Special Publication 35. p. 151–165. Google Scholar

    218.

    Lesquereux, L. 1876. Species of fossil marine plants from the Carboniferous measures: Geological Survey of Indiana Annual Report 7:134–145. Google Scholar

    219.

    Liddell, W. D., S. H. Wright, & C. E. Brett. 1997. Sequence stratigraphy and paleoecology of the Middle Cambrian Spence Shale in northern Utah and southern Idaho: Brigham Young Geologic Studies 42:59–78. Google Scholar

    220.

    Lin, J.-P., Y.-L. Zhao, I. A. Rahman, S. Xiao, & Y. Wang. 2010. Bioturbation in Burgess Shale-type Lagerstätten—Case study of trace fossil—body fossil association from the Kaili Biota (Cambrian Series 3), Guizhou, China: Palaeogeography, Palaeoclimatology, Palaeoecology 292(1–2):245–256. Google Scholar

    221.

    Linek, O. 1943. Die Buntsandstein-Kleinfährten von Nagold (Lingulichnulus nagoldensis n. g. n. sp., Merostomichnites triassicus n. sp.): Neues Jahrbuch für Mineralogie, Geologie und Paläontologie, Monatshefte Abteilung B:9–27. Google Scholar

    222.

    Linnaeus, C. 1758. Systema naturae per regna tria naturae scundum classes, ordines, genera, species, cum charateribus, differentiis sysnonymis, locis, 10th Edition, Tomus 1. Laurentii Salvii. Stockholm. 823 p. Google Scholar

    223.

    Linnarsson, J. G. O. 1869. Om några försteningar från Vestergötlands sandstenlager: Öfversikt av Kongliga Vetenskaps-Akademiens Handlingar 1869 3:337–357. Google Scholar

    224.

    Linnarsson, J. G. O. 1871. Geognostiska och paleontologiska iakttagelser öfver Eophytonsandstenen i Vestergötland: Kongliga Svenska Vetenskaps-Akademiens Handlingar 9(7): 1–19. Google Scholar

    225.

    Löffler, S.-B., & O. F. Geyer. 1994. Über Lebensspuren aus dem eozänen Belluno-Flysch (Nord-Italien): Paläontologische Zeitschrift 68(3–4):491–519. Google Scholar

    226.

    Loope, D. B., & L. Dingus. 1999. Mud-Filled Ophiomorpha from Upper Cretaceous Continental Redbeds of Southern Mongolia: An Ichnologic Clue to the Origin of Detrital, Grain-Coating Clays Palaios 14(5):451–458. Google Scholar

    227.

    Lorenz von Liburnau, J.R. 1902. Ergänzung sur Beschreibung der fossilen Halimeda juggeri: Akademie der Wissenschaften in Wien, Sitzungsberichte Mathmatisch-naturwissenschaftliche Klass 3(1):685–712, pl. 1–2. Google Scholar

    228.

    Lundgren, S. A. B. 1891. Studier öfver fossilförande lösa block: Geologiska Föreningen Stockholm Förhanlingar 12:111–121, text-fig. 1–2. Google Scholar

    229.

    Maas, O. 1902. Über Medusen aus dem Solenhofer Schiefer und der unteren Kreide der Karpathen: Palaeontographica 48:297–322. Google Scholar

    230.

    MacEachern, J. A., K. L. Bann, S. G. Pemberton, & M. K. Gingras. 2007a. The Ichnofacies Paradigm: High-Resolution Paleoenvironmental Interpretation of the Rock Record. In J. A. MacEachern, K. L. Bann, M. K. Gingras, & S. G. Pemberton, eds., Applied Ichnology: SEPM Short Course Notes 52. SEPM Tulsa, p. 27–64. Google Scholar

    231.

    MacEachern, J. A., S. G. Pemberton, K. L. Bann, & M. K. Gingras. 2007b. Departures from the Archetypal Ichnofacies: Effective Recognition of Physico-Chemical Stresses in the Rock Record. In J. A. MacEachern, K. L. Bann, M. K. Gingras, & S. G. Pemberton, eds., Applied Ichnology: SEPM Short Course Notes 52. SEPM Tulsa, p. 65–93. Google Scholar

    232.

    MacLeay, W. S. 1839. Notes on the Annelida. In R. J. Murchison, ed., The Silurian System, pt. 2. Murray. London, p. 699–701. Google Scholar

    233.

    MacNaughton, R. B. 2003. Planispiral burrows from a Recent lacustrine beach, Gander Lake Newfoundland: Canadian Field-Naturalist 117(4):577–581. Google Scholar

    234.

    MacNaughton, R. B., & G. M. Narbonne. 1999. Evolution and ecology of Neoproterozoic-Lower Cambrian trace fossils, NW Canada Palaios 14(2) :97–115. Google Scholar

    235.

    Macsotay, O. 1967. Huellas problematicas y su valor paleoecologico en Venezuela: Geos (Venezuela) 16:7–79. Google Scholar

    236.

    Mángano, M. G. 2011. Trace-fossil assemblages in a Burgess Shale-type deposit from the Stephen Formation at Stanley Glacier, Canadian Rocky Mountains: unraveling ecologic and evolutionary controls. In P. A. Johnston, & K. J. Johnston, eds., Proceedings of the International Conference on the Cambrian Explosion: Palaeontographica Canadiana. Canadian Society of Petroleum Geologists, Geological Association of Canada 31:89–107. Google Scholar

    237.

    Mángano, M. G., R. G. Bromley, D. A. T. Harper, A. T. Nielsen, M. P. Smith, & R. Vinther. 2012. Nonbiomineralized carapaces in Cambrian seafloor landscapes (Sirius Passet, Greenland): Opening a new window into early Phanerozoic benthic ecology: Geology 40(6):519–522. Google Scholar

    238.

    Mángano, M. G., L. A. Buatois, & F. Muñ Guinea. 2005. Ichnology of the Alfarcito Member (Santa Rosita Formation) of northwestern Argentina: animal-substrate interactions in a lower Paleozoic wavedominated shallow sea: Ameghiniana 42(4):641–668. Google Scholar

    239.

    Mángano, M. G., L. A. Buatois, R. Hofmann, O. Elicki, & R. Shinaq. 2013. Exploring the aftermath of the Cambrian explosion: The evolutionary significance of marginal- to shallow-marine ichnofaunas of Jordan: Palaeogeography, Palaeoclimatology, Palaeoecology 374:1–15. Google Scholar

    240.

    Mángano, M. G., L. A. Buatois, C. G. Maples, & R. R. West. 2000. A new ichnospecies of Nereites from Carboniferous tidal-flat facies of eastern Kansas, USA: implications for the Nereites—Neonereites debate: Journal of Paleontology 74(1): 149–157. Google Scholar

    241.

    Mángano, M. G., L. A. Buatois, & A. K. Rindsberg. 2002. Carboniferous Psammichnites: Systematic re-evaluation, taphonomy and autecology: Ichnos 9(1–2):1–22. Google Scholar

    242.

    Mángano, M. G., L. A. Buatois, R. R. West, & C. G. Maples. 2002. Ichnology of a Pennsylvanian equatorial tidal flat: the Stull shale member at Waverly, Eastern Kansas. Kansas Geological Survey (Bulletin 245): 133 p. Google Scholar

    243.

    Männil, R. M. 1966. O vertikalnykh norkakh zaryvaniya v Ordovikskikh izvestnyakakh Pribaltiki [A small vertically excavated cavity in Baltic Ordovician limestone]. In R. F. Hecker, ed., Organizm i sreda v geologischeskom proshlom. Paleontologicheskiy Institut. Akademiya Nauk SSSR. 1966. p. 200–207. Google Scholar

    244.

    Maples, C. G., & A. W. Archer. 1987. Redescription of early Pennsylvanian trace-fossil holotypes from the nonmarine Hindostan Whetstone beds of Indiana: Journal of Paleontology 61 (5):890–897. Google Scholar

    245.

    Maples, C. G., & R. R. West. 1989. Lockeia, not Pelecypodichnus: Journal of Paleontology 63(5):694–696. Google Scholar

    246.

    Marenco, K. N., & D. J. Bottjer. 2008. The importance of Planolites in the Cambrian substrate revolution: Palaeogeography, Palaeoclimatology, Palaeoecology 258(3): 189–199. Google Scholar

    247.

    Martin, A. J. 2013. Life Traces of the Georgia Coast. Revealing the Unseen Lives of Plants and Animals. Indiana University Press. Bloomington, Indiana. 670 p. Google Scholar

    248.

    Martin, K. D. 2004. A re-evaluation of the relationship between trace fossils and dysoxia. In D. Mcilroy, ed., The Application of Ichnology to Palaeoenvironmental and Stratigraphic Analysis. Geological Society. London. (Special Publication 228) p. 141–156. Google Scholar

    249.

    Matthew, G. F. 1891. Illustrations of the fauna of the St. John Group, no. V: Royal Society of Canada, Proceedings and Transactions 8(4):123–166, pl. 11–16. Google Scholar

    250.

    Matthew, G. F. 1910. Remarkable forms of the Little River Group. Proceedings and transactions of the Royal Society of Canada. Third series: 1909–1910. The Royal Society of Canada. Ottawa. 3(4):24 p. Google Scholar

    251.

    Maxey, G. B. 1958. Lower and Middle Cambrian stratigraphy in northern Utah and southeastern Idaho: Geological Society of America Bulletin 69(6):647–687. Google Scholar

    252.

    McCann, T., & R. K. Pickerill. 1988. Flysch trace fossils from the Cretaceous Kodiak Formation of Alaska: Journal of Paleontology 62(3):330–348. Google Scholar

    253.

    M‘Coy, F. 1850. On some genera and species of Silurian Radiata in the collection of the University of Cambridge: Annals and Magazine of Natural History (series 2) 6:270–290. Google Scholar

    254.

    M‘Coy, F. 1851. On some new Protozoic Annulata: Annals and Magazine of Natural History (series 2) 7:394–396. Google Scholar

    255.

    McIlroy, D., T. P. Crimes, & J. C. Pauley. 2005. Fossils and matgrounds from the Neoproterozoic Longmyndian Supergroup, Shropshire, UK: Geological Magazine 142(4):441–455. Google Scholar

    256.

    McMenamin, M. A. 1996. Ediacaran biota from Sonora, Mexico.Proceedings of the National Academy of Sciences, U.S.A. 93:4990–4993. Google Scholar

    257.

    Menon, L. R., D. McIlroy, A. G. Liu, & M. D. Brasier. 2015. The dynamic influence of microbial mats on sediments: fluid escape and pseudofossil formation in the Ediacaran Longmyndian Supergroup, UK: Journal of the Geological Society 173(1):177–185. Google Scholar

    258.

    Mikuláš, R. 1995. Trace fossils from the Paseky Shale (Early Cambrian, Czech Republic): Journal of the Czech Geological Society 40(4) :37–54. Google Scholar

    259.

    Miller, M. F., & S. E. Small. 1997. A semiquantitative field method for evaluating bioturbation on bedding planes Palaios 12:391–396. Google Scholar

    260.

    Miller, R. F. 1996. Location of trace fossils and problematica of George Frederic Matthew from Part W, Treatise on Invertebrate Paleontology: Journal of Paleontology 70(1): 169–171. Google Scholar

    261.

    Miller, S. A., & C. B. Dryer. 1878. Contributions to paleontology, no. 1: Journal of the Cincinnati Society of Natural History 1:24–39. Google Scholar

    262.

    Miller, S. A. 1889. North American Geology and Paleontology-for the use of Amateurs, Students and Scientists. 664 p. Google Scholar

    263.

    Minier, N. J., S. J. Braddy, & R. B. Davis. 2007. Between a rock and a hard place: arthropod trackways and ichnotaxonomy Lethaia 40(4):365–375. Google Scholar

    264.

    Minter, N. J., M. G. Mángano, & J.-B. Caron. 2012. Skimming the surface with Burgess Shale arthropod locomotion: Proceedings of the Royal Society, Biological Sciences 279(1):1613–1620. Google Scholar

    265.

    Morrissey, L. B., & S. J. Braddy. 2004. Terrestrial trace fossils from the Lower Old Red Sandstone, southwest Wales: Geological Journal 39(3–4):315–336. Google Scholar

    266.

    Morshedian, A., J. A. MacEachern, & S. E. Dashtgard. 2012. Integrated Ichnology, Sedimentology and Stratigraphy of the Lower Cretaceous Sparky Alloformation (Mannville Group), Lloydminster Area, westcentral Saskatchewan, Canada: Bulletin of Canadian Petroleum Geology 60(2):69–91. Google Scholar

    267.

    Mount, J. F. 1982. Storm-surge-ebb origin of hummocky cross-stratified units of the Andrews Mountain Member, Campito Formation (Lower Cambrian), White-Inyo Mountains, eastern California: Journal of Sedimentary Petrology 52(3):941–958. Google Scholar

    268.

    Murchison, R. I. 1839. The Silurian System. Murray, John. London. 768 p. Google Scholar

    269.

    Nara, M., & Y. Ikari. 2011. “Deep-sea bivalvian highways”: An ethological interpretation of branched Protovirgularia of the Palaeogene Muroto-Hanto Group, southwestern Japan: Palaeogeography, Palaeoclimatology, Palaeoecology 305(1):250–255. Google Scholar

    270.

    Narbonne, G. M., & J. D. Aitken. 1990. Ediacaran fossils from the Sekwi Brook area, Mackenzie mountains, northwestern Canada: Palaeontology 33(4):945–980. Google Scholar

    271.

    Narbonne, G. M., P. M. Myrow, E. Landing, & M. M. Anderson. 1987. A candidate stratotype for the Precambrian-Cambrian boundary, Fortune Head, Burin Peninsula, southeastern Newfoundland: Canadian Journal of Earth Sciences 24(7):1277–1293. Google Scholar

    272.

    Naruse, H., & K. Nifuku. 2008. Three-dimensional morphology of the ichnofossil Phycosiphon incertum and its implication for paleoslope inclination. Palaios 23(5):270–279. Google Scholar

    273.

    Nicholson, H. A. 1873. Contributions to the study of the errant annelides of the older Palaeozoic rocks: Royal Society of London Proceedings 21:288–290. Google Scholar

    274.

    Nicholson, H. A., & G. J. Hinde. 1875. Notes on the fossils of the Clinton, Niagara and Guelph formations of Ontario, with descriptions of new species: Canadian Journal of Sciences, Literature, and History, New Series 14:137–160 (actual pagination 137–152, 137–144). Google Scholar

    275.

    Oosterink, H. W., & H. Winkelhorst. 2013. Probable remains of jellyfish (Cnidaria, Scyphozoa) from the lower Middle Triassic (Anisian) of Winterswijk, eastern Netherlands: Netherlands Journal of Geosciences 92(1):61–67. Google Scholar

    276.

    Oriel, S. S., & F. C. Armstrong. 1971. Uppermost Precambrian and lowest Cambrian rocks in southeastern Idaho: U.S. Geological Survey Professional Papers 394:1–52. Google Scholar

    277.

    Orłowski, S. 1989. Trace fossils in the Lower Cambrian sequence in the Swietokrzyskie Mountains, Central Poland: Acta Palaeontology Polonica 34(3) :211–231. Google Scholar

    278.

    Orłowski, S. 1992. Trilobite trace fossils and their stratigraphical significance in the Cambrian sequence of the Holy Cross Mountains, Poland: Geological Journal 27:15–34. Google Scholar

    279.

    Orłowski, S., & A. Żylińska. 2002. Lower Cambrian trace fossils from the Holy Cross Mountains, Poland: Geological Quarterly 46(2): 135–146. Google Scholar

    280.

    Osgood, R. G. 1970. Trace fossils of the Cincinnati area: Palaeontographica Americana 6:277–444. Google Scholar

    281.

    Otto, E. v. 1854. Additamente zur Flora des Quadergebirges in Sachsen. G. Mayer. Leipzig (Part 2):53 p. Google Scholar

    282.

    Paczesna, J. 1996. The Vendian and Cambrian ichnocoenoses from the Polish part of the East-European Platform: Prace Państwowego Instytutu Geologiczny 153:1–77. Google Scholar

    283.

    Palmer, A. R. 1960. Some aspects of the early Upper Cambrian stratigraphy of White Pine County, Nevada and vicinity, Geology of east-central Nevada Intermountain Association of Petroleum Geologists Guidebook 11th Annual Field Conference, p. 53–58. Google Scholar

    284.

    Palmer, A. R., & D. P. Campbell. 1976. Biostratigraphic implications of trilobite biofacies, Albertella Zone, Middle Cambrian, western United States: Brigham Young University Geological Studies 23(2):39–50. Google Scholar

    285.

    Pandey, D. K., A. Uchman, V. Kumar, & R. S. Shekhawat. 2014. Cambrian trace fossils of the Cruziana ichnofacies from the Bikaner-Nagaur Basin, northwestern Indian Craton: Journal of Asian Earth Sciences 81:129–141. Google Scholar

    286.

    Paranjape, A. R., K. G. Kulkarni, & S. S. Gurav. 2013. Significance of Lockeia and associated trace fossils from the Bada Bagh Member, Jaisalmer Formation, Rajasthan: Indian Academy of Sciences Journal of Earth System Sciences 122(5):1359–1371. Google Scholar

    287.

    Peel, J. S. 2010. Articulated hyoliths and other fossils from the Sirius Passet Lagerstätte (early Cambrian) of North Greenland: Bulletin of Geosciences 85(3):385–394. Google Scholar

    288.

    Pemberton, S. G., & R. W. Frey. 1982. Trace fossil nomenclature and the Planolites-Palaeophycus dilemma: Journal of Paleontology 56(4):843–881. Google Scholar

    289.

    Pemberton, S. G., R. W. Frey, & R. G. Bromley. 1988. The ichnotaxonomy of Conostichus and other plug-shaped ichnofossüs: Canadian Journal of Earth Sciences 25(6):866–892. Google Scholar

    290.

    Pemberton, S. G., & B. Jones. 1988. Ichnology of the Pleistocene Ironshore Formation, Grand Cayman Island, British West Indies: Journal of Paleontology 62(4):495–505. Google Scholar

    291.

    Pemberton, S. G., M. Spila, A. J. Pulham, T. Saunders, J. A. MacEachern, D. Robbins, & I. K. Sinclair. 2001. Ichnology and sedimentology of shallow to marginal marine systems: Ben Nevis and Avalon reservoirs, Jeanne d'Arc Basin. Geological Association of Canada Short Course Notes 15. 343 p. Google Scholar

    292.

    Pemberton, S. G., & D. M. Wightman. 1992. Ichnological characteristics of brackish water deposits. In S. G. Pemberton, ed., Applications of Ichnology to Petroleum Exploration: A Core Workshop. SEPM Core Workshop Notes 17. p. 141–167. Google Scholar

    293.

    Peng, S., L. E. Babcock, & R. A. Cooper. 2012. The Cambrian Period. In F. M. Gradstein, J. G. Ogg, M. Schmitz, and G. Ogg, eds., The geologic time scale. 2:437–488. Google Scholar

    294.

    Péron, F., & C. A. Lesueur. 1810. Tableau des caractères génériques et spécifiques de toutes les espèces de méduses connues jusqu'à ce jour: Annales du Muséum d'Histoire Naturelle 14:325–366. Google Scholar

    295.

    Petrovich, R. 2001. Mechanisms of fossilization of the soft-bodied and lightly armored faunas of the Burgess shale and of some other classical localities: American Journal of Science 301:683–726. Google Scholar

    296.

    Pickerill, R. K. 1981. Trace fossils in a Lower Palaeozoic submarine canyon sequence—the Siegas Formation of northwestern New Brunswick, Canada: Atlantic Geology 17(1):37–58. Google Scholar

    297.

    Pickerill, R. K. 1991. The trace fossil Neonereites multiserialis Pickerill and Harland, 1988 from the Devonian Wapske Formation, northwest New Brunswick: Atlantic Geology 27:119–126. Google Scholar

    298.

    Pickerill, R. K. 1995. Deep-water marine Rusophycus and Cruziana from the Ordovician Lotbinière Formation of Quebec: Atlantic Geology 31(2):103–108. Google Scholar

    299.

    Pickerill, R. K., & L. R. Fyffe. 1999. The stratigraphic significance of trace fossils from the Lower Paleozoic Baskahegan Lake Formation near Woodstock, west-central New Brunswick: Atlantic Geology 35:205–214. Google Scholar

    300.

    Pickerill, R. K. & T. L. Harland. 1988. Trace fossils from Silurian slope deposits, North Greenland: Grønlands Geologiske Undersøgelse 137:119–133. Google Scholar

    301.

    Plaziat, J.-C., & M. Mahmoudi. 1988. Trace fossils attributed to burrowing echinoids: a revision including new ichnogenus and ichnospecies: Geobios 21(2):209–233. Google Scholar

    302.

    Plička, M., & Z. Siráňová. 1989. Hostynichnium duplex ichnogen. n. sp. n.—A new trace fossil from the Carpathian flysch of Czechoslovakia: Západné Karpaty: Séria Paleontología 13:109–112. Google Scholar

    303.

    Pollard, J. E. 1981. A comparison between the Triassic trace fossils of Cheshire and south Germany: Palaeontology 24(3):555–588. Google Scholar

    304.

    Pollard, J. E. 1985. Isopodichnus, related arthropod trace fossils and notostracans from Triassic fluvial sediments: Transactions of the Royal Society of Edinburgh: Earth Sciences 76(2–3):273–285. Google Scholar

    305.

    Prantl, F. 1945. Dve záhadné zkamenliny (stopy) z vrstev chrustenických–dδ2: Rozpravy II, Třidy Ceské Akademie 55(1):3–8. Google Scholar

    306.

    Prantl, F. 1946. Two new problematic trails from the Ordovician of Bohemia: Académie Techéque des Sciences, Bulletin International Classe des Sciences Mathématiques, Naturelles et da la Médecine 46(3):49–59. Google Scholar

    307.

    Quatrefages, M. A. d. 1849. Note sur la Scolicia prisca (A. de Q.), annélide fossile de la craie: Annales des Sciences Naturelles 3(12):265–266. Google Scholar

    308.

    Rajchel, J., & A. Uchman. 2012. Ichnology of Upper Cretaceous deepsea thick-bedded flysch sandstones: Lower Istebna Beds, Silesian Unit (Outer Carpathians, southern Poland): Geologica Carpathica 63(2): 107–120. Google Scholar

    309.

    Resser, C. E. 1939. The Spence Shale and its fauna (with six plates): Smithsonian Miscellaneous Collections 97(12): 1–37. Google Scholar

    310.

    Retrum, J. B., S. T. Hasiotis, & R. L. Kaesler. 2011. Neoichnological experiments with the freshwater ostracode Heterocypris incongruens: Implications for reconstructing aquatic settings Palaios 26(8): 509–518. Google Scholar

    311.

    Richter, R. 1850. Aus der thüringischen Grauwacke: Zeitschrift der Deutschen Geologischen Gesellschaft 2:198–206, pl. 8–9. [In German]. Google Scholar

    312.

    Richter, R. 1853. Gaea von Salfeld: Programm der Realschule Saalfeld, p. 3–32. Google Scholar

    313.

    Richter, R. 1871. Aus dem Thüringischen Schiefergebirge: Deutsche Geologische Gesellschaft Zeitschrift 23: 231–256. Google Scholar

    314.

    Richter, R. 1924. Flachseebeobachtungen zur Paläontologie und Geologie VII–XI: Senckenbergiana 6: 119–165. Google Scholar

    315.

    Richter, R. 1937. Marken und Spuren aus allen Zeiten, I–II: Senckenbergiana 19: 150–169. Google Scholar

    316.

    Rindsberg, A. K. 1994. Ichnology of the Upper Mississippian Hartselle Sandstone of Alabama, with Notes on Other Carboniferous Formations: Geological Survey of Alabama Bulletin 158: 1–107. Google Scholar

    317.

    Rindsberg, A. K., & D. C. Kopaska-Merkel. 2005. Treptichnus and Arenicolites from the Steven C. Minkin Paleozoic footprint site (Langsetrian, Alabama, USA): Pennsylvanian Footprints in the Black Warrior Basin of Alabama: Alabama Paleontological Society Monograph 1: 121–141. Google Scholar

    318.

    Robison, R. A. 1960. Lower and middle Cambrian stratigraphy of the eastern Great Basin, Geology of east-central Nevada Intermountain Association of Petroleum Geologists Guidebook 11th Annual Field Conference, p. 43–52. Google Scholar

    319.

    Robison, R.A. 1965. Middle Cambrian Eocrinoids from the Western North America: Journal of Paleontology 39(3):355–364. Google Scholar

    320.

    Robison, R. A. 1969. Annelids from the Middle Cambrian Spence Shale of Utah: Journal of Paleontology 43(5): 1169–1173. Google Scholar

    321.

    Robison, R. A. 1976. Middle Cambrian trilobite biostratigraphy of the Great Basin: Brigham Young University Geology Studies 23(2): 93–109. Google Scholar

    322.

    Robison, R. A. 1991. Middle Cambrian biotic diversity: examples from four Utah Lagerstätten. In A. M. Simonetta, & S. Conway Morris, eds., The early evolution of Metazoa and the significance of problematic taxa. Cambridge University Press. Cambridge, p. 77–98. Google Scholar

    323.

    Robison, R. A., & L. E. Babcock. 2011. Systematics, paleobiology, and taphonomy of some exceptionally preserved trilobites from Cambrian Lagerstätten of Utah: The University of Kansas Paleontological Contributions 5:1–47. Google Scholar

    324.

    Romano, M., & B. Melendez. 1985. An arthropod (merostome) ichnocoenosis from the Carboniferous of northwest Spain. In J. T. Dutro, & H. W. Pfefferkorn, eds., Neuvième Congres International de Stratigraphie et de Géologie du Carbonifère, Washington D.C. and Champaign—Urbana. Southern Illinois University Press. Carbondale and Edwardsville, Illinois. Compte Rendu 5:317–325. Google Scholar

    325.

    Sadlok, G. 2010. Trace fossil Cruziana tenella from the Furongian (Upper Cambrian) deposits of Poland: Acta Geologica Polonica 60(3) :349–355. Google Scholar

    326.

    Salter, J. W. 1857. On Annelide-burrows and Surface-markings from the Cambrian Rocks of the Longmynd. No. 2: Quarterly Journal of the Geological Society 13(1–2) : 199–206. Google Scholar

    327.

    Savage, N. M. 1971. A varvite ichnocoenosis from the Dwyka Series of Natal Lethaia 4(2):217–233. Google Scholar

    328.

    Savrda, C. E., A. D. Blanton-Hooks, J. W. Collier, R. A. Drake, R. L. Graves, A. G. Hall, A. I. Nelson, J. C. Slone, D. D. Williams, & H. A. Wood. 2000. Taenidium and associated ichnofossüs in fluvial deposits, Cretaceous Tuscaloosa Formation, eastern Alabama, southeastern USA: Ichnos 7(3):227–242. Google Scholar

    329.

    Schaffer, F. X. 1928. Homosiroidea florentina n.g., n. sp., ein Fucus aus der Kreide der Umgebung von Florenz: Paläontologische Zeitschrift 10:212–215. Google Scholar

    330.

    Schafhäutl, K. E. 1851. Geognostische Untersuchungen des Südbayerischen Alpengebirges. Literarisch-artistische Anstalt. München. 208 p. Google Scholar

    331.

    Schatz, E. R., M. G. Mángano, L. A. Buatois, & C. Oscar Limarino. 2011. Life in the Late Paleozoic Ice Age: Trace fossils from glacially influenced deposits in a Late Carboniferous fjord of western Argentina: Journal of Paleontology 85(3):502–518. Google Scholar

    332.

    Schembri, P. J., A. Deidun, & P. J. Vella. 2010. First record of Cassiopea andromeda (Scyphozoa: Rhizostomeae: Cassiopeidae) from the central Mediterranean Sea: Marine Biodiversity Records 3(e6):1–2. Google Scholar

    333.

    Schindewolf, O. H. 1921. Studien aus dem Marburger Buntsandstein I,II: Senckenbergiana 3:33–49. Google Scholar

    334.

    Schurf, M., A. Uchman, & M. Kümmel. 2001. Upper Triassic (Keuper) non-marine trace fossils from the HaBberge area (Franconia, southeastern Germany): Paläontologische Zeitschrift 75(1):71–96. Google Scholar

    335.

    Scott, J. J., & M. E. Smith. 2015. Trace Fossils of the Eocene Green River Lake Basins, Wyoming, Utah, and Colorado. In M. E. Smith, and A. R. Carroll, eds., Stratigraphy and Paleolimnology of the Green River Formation, Western USA. Springer. Netherlands, p. 313–350. Google Scholar

    336.

    Seilacher, A. 1953a. Studien zur Palichnologie 2. Die fossilen Ruhespuren (Cubichnia): Neues Jahrbuch für Geologie und Paläontologie, Abhandlungen 98:87–124. Google Scholar

    337.

    Seilacher, A. 1953b. Studien zur Palichnologie 1. über díe Methoden der Palichnologie: Neues Jahrbuch für Geologie und Paläontologie, Abhandlungen 96:421–452. Google Scholar

    338.

    Seilacher, A. 1955a. 5. Spuren und Fazies im Unterkambrium. In O. H. Schindewolf, and A. Seilacher, eds., Beiträge zur Kenntnis des Kambriums in der Salt Range (Pakistan). Akademie der Wissenschaften und der Literatur zu Mainz, mathematisch-naturwissenschaftliche Klasse, Abhandlungen 10:373–399. Google Scholar

    339.

    Seilacher, A. 1955b. 4. Spuren und Lebensweise der Trilobiten. In O. H. Schindewolf, & A. Seilacher, eds., Beiträge zur Kenntnis des Kambriums in der Salt Range (Pakistan). Akademie der Wissenschaften und der Literatur zu Mainz, mathematisch- naturwissenschftliche Klasse, Abhandlungen 10:342–372. Google Scholar

    340.

    Seilacher, A. 1960. Lebensspuren als Leitfossilien: Geologische Runschau 49:41–50. Google Scholar

    341.

    Seilacher, A. 1962. Paleontological Studies on Turbidite Sedimentation and Erosion: The Journal of Geology 70(2):227–234. Google Scholar

    342.

    Seilacher, A. 1970. Cruziana stratigraphy of “non-fossiliferous” Palaeozoic sandstones. In T. P. Crimes, and J. C. Harper, eds., Trace fossils. Geological Journal Special Publication 3:447–476. Google Scholar

    343.

    Seilacher, A. 1977. Pattern analysis on Paleodictyon and related trace fossils. In T. P. Crimes, & J. C. Harper, eds., Trace Fossils 2: Proceedings from the 25th International Geological Conference. Geological Journal Special Issue 9:289–334. Google Scholar

    344.

    Seilacher, A. 1985. Trilobite palaeobiology and substrate relationships: Transactions of the Royal Society of Edinburgh Earth Sciences 76(2–3):231–237. Google Scholar

    345.

    Seilacher, A. 1990. Paleozoic trace fossils. In R. Said, ed., The geology of Egypt. Balkema. Rotterdam, Brookfield, p. 649–670. Google Scholar

    346.

    Seilacher, A. 2007. Trace fossil analysis. Springer. Berlin Heidelberg. 226 p. Google Scholar

    347.

    Seilacher, A. T., & E. Seilacher. 1994. Bivalvian trace fossils: a lesson from actuopaleontology: Courier Forschungsinstitut Senckenberg 169:5–15. Google Scholar

    348.

    Selwyn, A. R. C. 1890. Tracks of organic origin in rocks of the Animikie Group: American Journal of Science 139:145–147. Google Scholar

    349.

    Serpagli, E. 2005. First record of the ichnofossil Atollites from the Late Cretaceous of the Northern Apennines, Italy: Acta Palaeontologica Polonica 50(2):4 03–408. Google Scholar

    350.

    Shah, S. K., & C. S. Sudan. 1983. Trace fossils from the Cambrian of Kashmir and their stratigraphic significance: Journal of the Geological Society of India 24(4): 194–202. Google Scholar

    351.

    Shone, R. W. 1978. Giant Cruziana from the Beaufort Group: Transactions of the Geological Society of South Africa 81:327–329. Google Scholar

    352.

    Shone, R. W. 1979. “Giant Cruziana from the Beaufort Group”: Transactions of the Geological Society of South Africa 82:371–375. Google Scholar

    353.

    Smith, A., S. J. Braddy, S. B. Marriott, & D. E. G. Briggs. 2003. Arthropod trackways from the Early Devonian of South Wales: a functional analysis of producers and their behaviour: Geological Magazine 140(1):63–72 Google Scholar

    354.

    Smith, J. J., & S. T. Hasiotis. 2008. Traces and burrowing behaviors of the cicada nymph Cicadetta calliope: Neoichnology and paleoecological significance of extant soil-dwelling insects Palaios 23(8):503–513. Google Scholar

    355.

    Smith, J. J., S. T. Hasiotis, M. J. Kraus, & D. T. Woody. 2008a. Relationship of floodplain ichnocoenoses to paleopedology, paleohydrology, and paleoclimate in the Wülwood Formation, Wyoming, during the Paleocene—Eocene Thermal Maximum Palaios 23(10):683–699. Google Scholar

    356.

    Smith, J. J., S. T. Hasiotis, M. J. Kraus, & D. T. Woody. 2008b. Naktodemasis bowni: new ichnogenus and ichnospecies for adhesive meniscate burrows (AMB), and paleoenvironmental implications, Paleogene Wülwood Formation, Bighorn Basin, Wyoming: Journal of Paleontology 82(2):267–278. Google Scholar

    357.

    Smith, J. J., S. T. Hasiotis, M. J. Kraus, & D. T. Woody. 2009. Transient dwarfism of soil fauna during the Paleocene—Eocene Thermal Maximum: Proceedings of the National Academy of Sciences 106(42):17655–17660. Google Scholar

    358.

    Sowerby, J. 1829. The mineral conchology of Great Britain. Richard Taylor. London. 6:230 p. Google Scholar

    359.

    Sprinkle, J., & D. Collins. 2006. New eocrinoids from the Burgess Shale, southern British Columbia, Canada, and the Spence Shale, northern Utah, USA: Canadian Journal of Earth Sciences 43:303–322. Google Scholar

    360.

    Sternberg, G. K. v. 1833. Versuch einer geognostisch-botanischen Darstellung der Flora der Vorwelt. Parts 5–6. Fleischer. Prague. 80 p. Google Scholar

    361.

    Stachacz, M. 2012. Ichnology of Czarna Shale Formation (Cambrian, Holy Cross Mountains, Poland): Annales Societatis Geologo rum Poloniae 82:105–120. Google Scholar

    362.

    Stachacz, M. 2016. Ichnology of the Cambrian Odesêki Sandstone Formation (Holy Cross Mountains, Poland): Annales Societatis Geologorum Poloniae 86:291–328. Google Scholar

    363.

    Strezeboński, P., & A. Uchman. 2015. The trace fossil Gyrophyllites Ín deep-sea siliciclastic deposits of the Istebna Formation (Upper Cretaceous—Palaeocene) of the Carpathians: An example of biologically controlled distribution: Palaeogeography, Palaeoclimatology, Palaeoecology 426:260–274. Google Scholar

    364.

    Tarhan, L. G., S. Jensen, & M. L. Droser. 2011. Furrows and firmgrounds: evidence for predation and implications for Palaeozoic substrate evolution in Rusophycus burrows from the Silurian of New York: Lethaia 45(3):329–34l. Google Scholar

    365.

    Torell, O. M. 1870. Petrificata Suecana Formationis Cambricae: Lunds Universitets Arsskrift 6:1–14. Google Scholar

    366.

    Trewin, N. H. 1976. Isopodichnus in a trace fossil assemblage from the Old Red Sandstone: Lethaia 9:29–37. Google Scholar

    367.

    Trewin, N. H. 1994. A draft system for the identification and description of arthropod trackways: Palaeontology 37(4) :811–823. Google Scholar

    368.

    Trewin, N. H., & K. J. McNamara. 1995. Arthropods invade the land: trace fossils and palaeoenvironments of the Tumblagooda Sandstone (?late Silurian) of Kalbarri, Western Australia: Transactions of the Royal Society of Edinburgh Earth Sciences 85(3):177–210. Google Scholar

    369.

    Twenhofel, W. H. 1927. Geology of Anticosti Island. Canada Department of Mines Geological Survey. Memoir 154:481 p. Google Scholar

    370.

    U. S. National Institutes of Health, 1997–2015, ImageJ: Bethesda, Maryland, USA.  http://imagej.nih.gov/ij/Google Scholar

    371.

    Ubaghs, G., & R. A. Robison. 1985. A new homoiostelean and a new eocrinoid from the Middle Cambrian of Utah: University of Kansas Paleontological Contributions 115:1–24. Google Scholar

    372.

    Uchman, A. 1991. Trace fossils of the Inoceramian beds and the Szczawnica Formation in the Krynica and Bystrzyca Zones of the Magura Nappe: Przeglad Geologiczny 39(4):207–212 (in Polish with English summary). Google Scholar

    373.

    Uchman, A. 1992. Trace fossils of the Eocene thin- and medium-bedded flysch of the Bystrica Zone of the Magura Nappe in Poland: Przeglad Geologiczny 40(7):430–435 (in Polish with English summary). Google Scholar

    374.

    Uchman, A. 1995. Taxonomy and palaeoecology of flysch trace fossils: The Mamoso-Arenacea Formation and associated facies (Miocene, Northern Apennines, Italy): Beringeria 15:3–115. Google Scholar

    375.

    Uchman, A. 1998. Taxonomy and ethology of flysch trace fossils: a revision of the Marian Ksiązkiewicz collection and studies of complementary material: Annales Societatis Geologorum Poloniae 68:105–218. Google Scholar

    376.

    Uchman, A. 1999. Ichnology of the Rhenodanubian Flysch (Lower Cretaceous—Eocene) in Austria and Germany: Beringeria 25: 67–173. Google Scholar

    377.

    Uchman, A. 2001. Eocene flysch trace fossils from the Hecho Group of the Pyrenees, northern Spain: Beringeria 28:3–41, 43 fig., 14 pl. Google Scholar

    378.

    Uchman, A. 2007. Trace fossils of the Pagliaro Formation (Paleocene) ín the North Apennines, Italy: Beringeria 37:217–237, 3 text fig., 2 tables, 4 pl. Google Scholar

    379.

    Uchman, A. 2008a. Cretaceous—Neogene flysch deposits of the Outer Carpathians. In A. Uchman, ed., Types of invertebrate trace fossils from Poland: an illustrated catalogue. Polish Geological Institute. Warszawa. Poland, p. 24–65. Google Scholar

    380.

    Uchman, A. 2008b. Stop 8—Zbludza—Beloveža Formation (Eocene) and Bystrica Formation (Eocene): outer fan ichnology and sequential colonization of turbkikes. In G. Pieńkowski, & A. Uchman, eds., Ichnological Sites of Poland: The Holy Cross Mountains and the Carpathian Flysch—The Second International Congress on Ichnology, Cracow, Poland, August 29- September 8, 2008; Pre-Congress and Post-Congress Field Trip Guidebook. Polish Geological Institute. Warszawa, p. 124–131. Google Scholar

    381.

    Uchman, A., R. G. Bromley, & S. Leszczyňski. 1998. Ichnogenus Treptichnus in Eocene flysch, Carpathians, Poland: Taxonomy and preservation: Ichnos 5(4):269–275. Google Scholar

    382.

    Uchman, A., D. Drygant, M. Paszkowski, S. J. Porçbski, & E. Turnau. 2004. Early Devonian trace fossils in marine to non-marine redbeds in Podolia, Ukraine: palaeoenvkonmental implications: Palaeogeography, Palaeoclimatology, Palaeoecology 214(1):67–83. Google Scholar

    383.

    Uchman, A., V. Kazakauskas, & A. Gaigalas. 2009. Trace fossils from Late Pleistocene varved lacustrine sediments in eastern Lithuania: Palaeogeography, Palaeoclimatology, Palaeoecology 272(3):199–211. Google Scholar

    384.

    Uchman, A., R. Mikuláš, & A. K. Rindsberg. 2011. Mollusc trace fossils Ptychoplasma Fenton and Fenton, 1937 and Oravaichnium Plíčka and Uhrová, 1990: Their type material and ichnospecies Geobios 44(4):387–397. Google Scholar

    385.

    Vannier, J., I. Calandra, C. Gaillard, & A. Żylińska. 2010. Priapulid worms: Pioneer horizontal burrowers at the Precambrian-Cambrian boundary Geology 38(8):711–714. Google Scholar

    386.

    Vassoevich, N. B. 1932. O nekotorykh priznakakh pozvolyayushchikh otlicht'oprokinutoe polozhenie flishevykh obrazovaniy ot normalynogo: Akademie Nauka SSSR,Geologiduskiy Institut Trudy 2:47–64. Google Scholar

    387.

    Vassoevich, N. B. 1951. Usloviya obrazovaniya flisha [The conditions of the formation of flysch]. Gostoptekhizdat. Leningrad. 240 p. Google Scholar

    388.

    Vialov, O. S. 1962. Problematica of the Beacon Sandstone at Beacon Heights, West Antarctica: New Zealand Journal of Geology and Geophysics 5:718–732. Google Scholar

    389.

    Vialov, O. S. 1964. Zvezdchatye ieroglify iz Triasa severovostoka Sibiri [Star-shaped hieroglyphs from the Triassic of northeastern Siberia]: Akademija Nauk SSSR, Geologii i Geofiziki, Sibkskoe Otdelenie 5:112–115. Google Scholar

    390.

    Waggoner, B. & J. W. Hagadorn. 2002. New fossils from terminal Neoproterozoic strata of Southern Nye County, Nevada. In F. A. Corsetti, ed., Proterozoic—Cambrian of the Great Basin and Beyond: Volume and Guidebook for the Society for Sedimentary Geology SEPM 93:87–96. Google Scholar

    391.

    Walcott, C. D. 1890. Descriptive notes of new genera and species from the Lower Cambrian or Olenellus Zone of North America: United States National Museum Proceedings 12:33–46. Google Scholar

    392.

    Walcott, C. D. 1896. Fossil jelly fishes from the Middle Cambrian terrane: Proceedings of the United States National Museum 42:611–614. Google Scholar

    393.

    Walcott, C. D. 1908. Cambrian Geology and Paleontology: Cambrian Sections of the Cordilleran Area with Ten Plates: Smithsonian Miscellaneous Collections 53(5): 167–230. Google Scholar

    394.

    Walter, H., & B. Gaitzsch. 1988. Zur Ichnologie limnisch-terrestrischer Sedimentationsräume, Teil 2: Diplichnites minimus n. ichnosp. aus dem Permosiles des Flechtinger Höhenzuges: Freiberger Forschungsheft 100:73–84. Google Scholar

    395.

    Walter, M. R., R. Elphinstone, & G. R. Heys. 1989. Proterozoic and Early Cambrian trace fossils from the Amadeus and Georgina Basins, central Australia: Alcheringa 13(3):209–256. Google Scholar

    396.

    Wang, Y., J.-P. Lin, Y.-L. Zhao, & P. J. Orr. 2009. Palaeoecology of the trace fossil Gordia and its interaction with nonmineralizing taxa from the early Middle Cambrian Kaili Biota, Guizhou province, South China: Palaeogeography, Palaeoclimatology, Palaeoecology 277 1–2):141–148. Google Scholar

    397.

    Wang, Y., Y. L. Zhao, J. P. Lin, & P. L. Wang. 2004. Relationship between trace fossil Gordia and medusiform fossils Pararotadiscus from the Kaili Biota, Taijiang, Guizhou, and its significance: Geological Review 30(1):113–119. Google Scholar

    398.

    Webby, B. D. 1983. Lower Ordovician arthropod trace fossils from western New South Wales: Proceedings of the Linnean Society of New South Wales 107:39–74. Google Scholar

    399.

    Wells, A. T., D. J. Forman, L. C. Ranford, & P. J. Cook. 1970. Geology of the Amadeux basin, central Australia. Bureau of Mineral Resources Geology & Geophysics Bulletin 100:216 p. Google Scholar

    400.

    Weller, S. 1899. Kinderhook Journal Studies 1: The Fauna of the vermicular sandstone at Northview, Webster County, Missouri: Academy of Science of St. Louis Transactions 9:9–51, pl. 2–6. Google Scholar

    401.

    Wetzel, A., & R. G. Bromley. 1994. Phycosiphon incertum revisited: Anconichnus horizontalis is its junior subjective synonym: Journal of Paleontology 68(6): 1396–1402. Google Scholar

    402.

    Wetzel, A., & A. Uchman. 1997. Ichnology of deep-sea fan overbank deposits of the Ganei slates (Eocene, Switzerland)—a classical flysch trace fossil locality studied first by Oswald Heer: Ichnos 5(2) :139–162. Google Scholar

    403.

    Wetzel, A., & A. Uchman. 2001. Sequential colonization of muddy turbidites in the Eocene Beloveža Formation, Carpathians, Poland: Palaeogeography, Palaeoclimatology, Palaeoecology 168(1): 171–186. Google Scholar

    404.

    White, D. 1929. Flora of the Hermit Shale, Grand Canyon, Arizona. Carnegie Institute of Washington Publication 405:221 p. Google Scholar

    405.

    Willoughby, R. H., & R. A. Robison. 1979. Medusoids from the Middle Cambrian of Utah: Journal of Paleontology 53(2):494–500. Google Scholar

    406.

    Wilson, J. P., J. P. Grotzinger, W. W. Fischer, K. P. Hand, S. Jensen, A. H. Knoll, J. Abelson, J. M. Metz, N. McLoughlin, P. A. Cohen, & M.M. Tice. 2012. Deep-water incised valley deposits at the Ediacaran-Cambrian Boundary in southern Namibia contain abundant Treptichnus pedum Palaios 27(4):252–273. Google Scholar

    407.

    Yang, R. D., & Y. L. Zhao. 1999. Discovery on trace fossils from the Early—Middle Cambrian Kaili Formation of Taijiang, Guizhou: Acta Palaeontologica Sinica 38(1):58–65. Google Scholar

    408.

    Yang, S.-P. 1984. Silurian trace fossils from the Yangzi Gorges and their significance to depositional environments: Acta Palaeontologica Sinica 23:705–714. Google Scholar

    409.

    Yang, S.-P. 1994. Trace fossils from the Early—Middle Cambrian Kaili Formation in Taijang, Guizhou: Acta Palaeontologica Sinica 33(1):350–358. Google Scholar

    410.

    Yang, S., C. Hu, & Y. Sun. 1987. Discovery of Late Devonian trace fossils from Guodingshan district, Hanyang, China and its significance: Earth Science Journal of the Wuhan College of Geology 12:1–8. Google Scholar

    411.

    Yang, S.-P., & X.-C. Wang. 1991. Middle Cambrian Hsuchuangian trace fossils from southern North China Platform and their sedimentological significance: Acta Palaeontologica Sinica 30(1):74–89. Google Scholar

    412.

    Yang, Z., J. Yin, & T. He. 1982. Early Cambrian trace fossils from the Emei-Ganluo region, Sichuan, and other localities: Geological Review 28:291–298. Google Scholar

    413.

    Yochelson, E. L., & M. A. Fedonkin. 1997. The type specimens (Middle Cambrian) of the trace fossil Archaeonassa Fenton and Fenton: Canadian Journal of Earth Sciences 34(9): 1210–1219. Google Scholar

    414.

    Yochelson, E. L., & D. E. Schindel. 1978. A re-examination of Pennsylvanian trace fossil Olivellites: Journal of Research of the US Geological Survey 6(6):789–796. Google Scholar

    415.

    Young, F. G. 1972. Early Cambrian and older trace fossils from the Southern Cordillera of Canada: Canadian Journal of Earth Science 9:1–17. Google Scholar

    416.

    Zenker, J.C. 1836. Historisch-topographisches Taschenbuch von Jena und seiner Umgebung besonders in naturwissenschaftlicher und medicinischer Beziehung. Frommann. Jena. 338 p. Google Scholar

    417.

    Zhang, X.-G., J. Bergström, R. G. Bromley, & X.-G. Hou. 2007. Diminutive trace fossils in the Chengjianp; Lagerstätte: Terra Nova 19(1) :407–412. Google Scholar

    418.

    Zhang, X., & D. Wang. 1996. A restudy of Silurian—Devonian ichnofossils from northwestern Hunan area: Acta Palaeontologica Sinica 35(475–489). Google Scholar

    419.

    Zonneveld, J. P., M. K. Gingras, & T. W. Beatty. 2010. Diverse ichnofossil assemblages following the P-T mass extinction, Lower Triassic, Alberta and British Columbia, Canada: Evidence for shallow marine refugia on the northwestern coast of Pangaea Palaios 25(6): 368–392. Google Scholar

    420.

    Zonneveld, J.-P., S. G. Pemberton, T. D. A. Saunders, & R. K. Pickerill. 2002. Large, Robust Cruziana from the Middle Triassic of Northeastern British Columbia: Ethologic, Biostratigraphic,and Paleobiologic Significance Palaios 17(5):435–448. Google Scholar
    © 2018, The University of Kansas, Paleontological Institute
    Sean R. Hammersburg, Stephen T. Hasiotis, and Richard A. Robison "Ichnotaxonomy of the Cambrian Spence Shale Member of the Langston Formation, Wellsville Mountains, Northern Utah, Usa," Paleontological Contributions 2018(20), 1-66, (18 May 2018). https://doi.org/10.17161/1808.26428
    Published: 18 May 2018
    KEYWORDS
    Archaeonassa
    Cruziana
    Gyrophyllites
    ichnofossil
    Rusophycus
    Back to Top