You are on page 1of 15

Precambrian Research 399 (2023) 107227

Contents lists available at ScienceDirect

Precambrian Research
journal homepage: www.elsevier.com/locate/precamres

First record of trace fossils in the Ediacaran–Cambrian transition on the


northern Gondwana platform (Anti-Atlas, Morocco)
Abdelfattah Azizi a, *, Asmaa El Bakhouch a, Abderrazak El Albani b, Kalle Kirsimäe c,
Mouhssin El Halim a, Khadija El Hariri a, Mohamed Erragragui a, Ahmid Hafid a, Olev Vinn c
a
Laboratoire de Géo-ressources, Géoenvironnement et Génie civil, Département des Sciences de la Terre, Faculté des Sciences et Techniques, Université Cadi-Ayyad, BP
549, 40000 Marrakesh, Morocco
b
Université de Poitiers, UMR-CNRS, IC2MP 7285, 86022 Poitiers, France
c
Department of Geology, Institute of Ecology and Earth Sciences, University of Tartu, Ravila 14A, 50411 Tartu, Estonia

A R T I C L E I N F O A B S T R A C T

Keywords: The Anti-Atlas Mountains of Morocco preserve one of the most complete latest Neoproterozoic to Cambrian
Ediacaran stratigraphic successions worldwide, with high-resolution chemostratigraphic δ13Ccarb coverage. However, the
Cambrian exact stratigraphic position of the Ediacaran–Cambrian boundary remains unresolved. Until now, no trace fossils
Trace fossils
or body fossils have been recorded from the terminal Ediacaran strata in the Anti-Atlas Mountains. The only
Tabia Member
exception is a discoidal body fossil-like structure from the shallow-marine sandstones of the Ediacaran Tabia
Member, Adoudou Formation, in the Taroudant Group in the Igherm inlier. For the first time, we document the
occurrence of abundant trace fossils from the base of the Adoudou Formation, including Treptichnids, Mono­
morphichnus isp., Helminthoidichnites isp., Gordia isp., Palaeophycus isp., Planolites isp., and Conichnus isp. The
integration of both trace fossils and carbon isotope δ13Ccarb records suggests that the Ediacaran–Cambrian
transition in the Sous Basin can be placed at the contact between the Tifnout and Tabia Members, which che­
mostratigraphically coincides with the large negative δ13Ccarb excursion that had been previously suggested for
the Ediacaran–Cambrian boundary (BACE) in the western Anti-Atlas Mountains.

1. Introduction diversification of complex animals capable of burrowing and pene­


trating the sediment, thereby producing a dramatic increase in the depth
The Ediacaran–Cambrian boundary is one of the most interesting and intensity of bioturbation, known as the ‘Agronomic Revolution’ (e.
transitions in the history of life, characterized by the first mass extinc­ g., Seilacher, 1999; Bottjer et al., 2000; Seilacher et al., 2005; Buatois
tion of the Ediacaran biota in the fossil record (Darroch et al., 2015, et al., 2014, 2020). This bioturbation event was marked by a sudden
2018; Smith et al., 2017; Nance et al., 2022), the first appearance of shift in the trace fossil style from horizontal surficial traces in the late
diverse metazoans with biomineralized skeletons (Germs, 1995), and an Ediacaran to vertically penetrative ichnofabrics in the Cambrian. The
exceptional development of metazoan complexity and behaviour Ediacaran–Cambrian boundary is defined by the first appearance of the
(Erwin, 1999; Knoll and Carroll, 1999; Narbonne, 2004, 2005). How­ trace fossil Treptichnus pedum (Brasier et al., 1994). Secondary markers
ever, the exact timing of this bioevent and the stratigraphic position of include the last occurrences of Ediacaran-type body fossils and the first
the Ediacaran–Cambrian boundary remain poorly resolved. Reaching appearance of small shelly fossils (SSFs), as well as a large negative
this goal is complicated due to an approximately 20-million-year gap carbon isotope excursion, termed the BACE (BAsal Cambrian carbon
between the youngest preservation of Ediacaran body fossil assemblages isotope Excursion) (e.g., Kimura and Watanabe, 2001; Amthor et al.,
and that of Cambrian Stage 3 (Buatois et al., 2014). In contrast, the trace 2003; Smith et al., 2016; Nelson et al., 2022, 2023).
fossil record across the Ediacaran–Cambrian transition is far more In Morocco, the Ediacaran–Cambrian Adoudou Formation of the
diverse and continuous than that of the body fossils, providing valuable Taroudant Group is well exposed across the Anti-Atlas Mountains and
information about this critical biological transition. One of the ecolog­ serves as a standard for the global chemostratigraphic carbonate δ13C
ical manifestations of the Cambrian Explosion was a rapid spread and curve of the Ediacaran–Cambrian boundary interval (Maloof et al.,

* Corresponding author.
E-mail address: a.azizi@uca.ma (A. Azizi).

https://doi.org/10.1016/j.precamres.2023.107227
Received 11 April 2023; Received in revised form 22 September 2023; Accepted 2 November 2023
Available online 11 November 2023
0301-9268/© 2023 Elsevier B.V. All rights reserved.
A. Azizi et al. Precambrian Research 399 (2023) 107227

Fig. 1. A, simplified map of Africa showing the location of the Anti-Atlas on the northern border of the West African Craton (WAC); B, geological map of the Anti-
Atlas with locations of measured sections (redrawn and modified after Saadi et al., 1985); C, Composite section of the Taroudant and Tata Groups in the western Anti-
Atlas (δ13Ccarb data are from Maloof et al., 2005).

2005, 2010). Based on carbon isotope δ13Ccarb correlations, a negative by using an integration of the newly available ichnological data and
δ13Ccarb excursion (down to δ13Ccarb values of − 6 ‰) located at the carbonate δ13C chemostratigraphy.
bottom of the Tifnout Member (Tamjout bed) in the sous basin was
suggested to represent the Ediacaran–Cambrian boundary (BACE) 2. Geological setting and stratigraphy
(Maloof et al., 2005, 2010). However, neither body fossils nor trace
fossils have yet been found, except for a discoidal fossil preserved in The Anti-Atlas Mountains are an ~1300 km long NE-SW-trending
shallow-marine sandstones from the Tabia Member in the Igherm inlier belt on the northern edge of the Eburnian (c. 2 Ga) West African
that was originally interpreted as a fossil jellyfish (Houzay, 1979) and Craton (WAC) (Fig. 1A). This orogen is separated from the High Atlas
redescribed as Aspidella discs (Maloof et al., 2005; Letsch et al., 2019). Mountains to the north by the South Atlas Faults (Fig. 1B). The southern
The aims of this study are to (1) document the Adoudou Formation slope of the Anti-Atlas Mountains is composed of a Palaeozoic sedi­
ichnofauna and (2) discuss the most plausible position for the Ediacar­ mentary succession that overlies the Precambrian rocks cropping out
an–Cambrian boundary in the western Anti-Atlas Mountains of Morocco within numerous basement inliers (Fig. 1B), providing well-exposed

2
A. Azizi et al. Precambrian Research 399 (2023) 107227

Fig. 2. A lithostratigraphic section of the studied outcrop in the Zaouia locality, with associated carbonate oxygen and carbon isotope data, is presented. The
horizontal dashed line marks the Ediacaran–Cambrian boundary based on the lowest occurrence of the Fortunian trace fossils as well as the nadir of a negative
carbonate carbon isotope excursion that may correlate to the BACE (Corsetti and Hagadorn, 2000).

outcrops of crystalline and sedimentary rocks of Palaeo-, Meso- and which yield U–Pb baddeleyite ages of ~1710 Ma (Ikenne et al., 2017;
Neoproterozoic age (e.g., Thomas et al., 2002, 2004; Walsh et al., 2012; Youbi et al., 2013).
Hefferan et al., 2014; Ikenne et al., 2017). The Precambrian basement Relics of coeval oceanic crust rocks are preserved along the AAMF,
and its Palaeozoic cover were deformed in a thick-skinned fashion from the Siroua to Bou Azzer inliers (Fig. 1B), the Tachdamt Bleïda series
during the Hercynian orogeny (Faik et al., 2001; Baidder et al., 2016). and the associated ophiolitic complex (referred to as PII in the regional
The Palaeoproterozoic basement (referred to as PI in the regional literature, e.g., Choubert, 1963). The PII units document the Neo­
literature), occurring on the southwestern side of the Anti-Atlas Major proterozoic Pan-African cycle (c. 760–600 Ma), with the opening and
Fault (AAMF; Choubert, 1947) (Fig. 1B), is composed of schists, para­ closing of a Cryogenian ocean basin between the WAC’s northern rifted
gneisses, and migmatites that are intruded by batholiths yielding U–Pb margin and a northern continental block. The closure of the basin was
zircon ages ranging from ~2.2 to 2.0 Ga (e.g., Thomas et al., 2002; followed by widespread intracontinental deformation and continental
Walsh et al., 2002; Gasquet et al., 2008; O’Connor, 2010). The quartzite and marine sedimentation (Gasquet et al., 2008; El Hadi et al., 2010;
and limestone series of Taghdout and Jbel Lkest that overlie the Palae­ Blein et al., 2014; Hefferan et al., 2014; Soulaimani et al., 2018; Tri­
oproterozoic rocks were attributed to the Cryogenian, based on the Rb/ antafyllou et al., 2016).
Sr 789 ± 10 Ma age of the contact-metamorphosed wall-rocks of mafic Both PI and PII are unconformably covered by the Ouarzazate Su­
dykes (Clauer, 1974). Recent studies have highlighted the existence of pergroup (sensu Thomas et al., 2004) or PIII (Choubert, 1963). This unit
Mesoproterozoic mafic dikes within the Taghdout sedimentary series, consists of a ~2000 m-thick pile of subaerial, volcano-sedimentary

3
A. Azizi et al. Precambrian Research 399 (2023) 107227

series deposited between ~615 and 538 Ma within alluvial and fluvial
environments (e.g., Thomas et al., 2002; Gasquet et al., 2008; Walsh
et al., 2012;). The youngest U-Pb zircon age (538 ± 6 Ma) has been
reported from the Jbel Saghro inlier in the eastern Anti-Atlas Mountains
(Baidada et al., 2018), whereas the maximum depositional age of ~541
Ma has been reported from the Bas-Draa inlier in the western Anti-Atlas
Mountains and shifts the Precambrian–Cambrian boundary down from
the Lower Adoudou Formation, as proposed by previous studies, to the
uppermost Ouarzazate Supergroup or to the hiatus between both units
(Karaoui et al., 2015). PIII rocks are, in turn, covered by a thick suc­
cession of nonmetamorphic Palaeozoic successions (Fig. 1B and C). The
Cambrian sedimentary rocks that unconformably overlie the Ouarzazate
Supergroup are dominated by shallow-marine carbonate and siliciclastic
rocks and are subdivided into two main groups (Fig. 1C). The ~1000 m
thick Taroudant Group is subdivided into two formations: the Adoudou
and Taliwine Formations (Choubert, 1952). The Adoudou Formation is
subdivided into the Tabia and Tifnout members (sensu, Maloof et al.,
2005). The Tabia Member (la série de base, sensu Choubert, 1952) is
represented by a mixed siliciclastic–carbonate succession that is sub­
divided into three units (Benssaou and Hamoumi, 2001, 2004). The
‘basal conglomerate’ unit is composed of coarse-grained sandstone and Fig. 3. Cross-plot and linear correlation cart of the δ13C and δ18O values of the
conglomerate deposited in a marine fan-delta setting dominated by al­ 45 samples from the Zaouia section.
luvial processes; the middle ‘basal limestone’ unit preserves well-
developed cyanobacterial stromatolites formed in a shallow carbonate
4. Carbonate carbon isotope systems
platform, and the overlying ‘basal siltstone’ (studied in this paper;
Fig. 1D, 2) consists of alternating transgressive–regressive depositional
Carbonate δ13C and δ18O values exhibit rather large variations
sequences, displaying typical wave-generated structures, such as hum­
(Fig. 3, Table 1). The carbonate δ13C and δ18O cross-plot does not show a
mocky cross-stratification and polygonal ripples that record episodic
distinctive positive correlation between C and O isotopic values that
storm events (Benssaou and Hamoumi, 2001, 2004). The Tifnout
would indicate diagenetic alteration (Banner and Hanson, 1990; Kauf­
Member is dominated by stromatolitic carbonate deposited in peritidal
man and Knoll, 1995), and the studied rocks can be considered well
platforms. The Adoudou Formation is overlain by the regressive deposits
preserved (but see below).
of the Taliwine Formation, consisting of fine-grained siliciclastic sedi­
The basal limestone unit shows overall higher δ13Ccarb values, typi­
ments with several carbonate interbeds. The overlying fossiliferous Tata
cally varying between − 1 and 1 ‰, but they show large variations from
Group (Cambrian stage 2) is well exposed in the Sous Basin (~1000 m
− 2.6 to 2.1 ‰ (Fig. 2) over a few metres at the contact between the basal
thick) and is dominated by variegated shale with interbedded black
limestone and overlying basal siltite, which most likely indicates a
oolitic limestone. The Tata Group records a sudden shift in the compo­
diagenetic imprint in this interval. In the Tifnout Member, the δ13Ccarb
nents of ecosystems, marked by the disappearance of microbial con­
values range narrowly between − 2.9 and 3.8 ‰ and show a slight trend
sortia (stromatolite-dominated) and the emergence of thrombolites and
from values closer to − 4 ‰ in the lower part of the unit to values ca. − 3
shelly metazoans, including archaeocyaths, trilobites, chancelloriids,
‰ in the middle and upper parts of the studied section (Fig. 2).
hyoliths and calcimicrobes (e.g., Hupé, 1960; Schmitt and Monninger,
1977; Sdzuy, 1978; Schmitt, 1979; Debrenne and Debrenne, 1995;
5. Synopsis of trace fossils
Álvaro and Clausen, 2006; Álvaro and Debrenne, 2010; Clausen et al.,
2014; Azizi et al. 2022).
Trace fossils from the Tabia Member were divided into four cate­
gories of traces: simple horizontal scratches and trails; simple actively
3. Materials and methods
filled (massive) horizontal to oblique structures; horizontal burrows
with horizontal to vertical branches and vertical plug-shaped burrows. A
The Tabia Member is well exposed along a road crossing the Alma
wide variety of ichnofossils are associated with microbially induced
inlier (Fig. 1). One representative section (Zaouia section; GPS: N
sedimentary structures (MISSs) at the tops of siltstone and sandstone
29◦56′27′′, W 8◦4′42′′), comprising the ‘Basal siltite’ unit, was logged
beds. Most of the beds containing MISSs yield well-preserved trace
(Fig. 2), and trace fossil samples were collected during field campaigns
fossils. Trace fossils are variably distributed on exposed bedding
in 2018, 2019 and 2022. Photographs were taken with a Canon EOS
surfaces.
650D. Polished slabs were scanned by a digital scanner (EPSON
EXPRESSION 10000 XL). Specimens are housed at the Faculté des Sci­
ences et Techniques, Université Cadi-Ayyad. 5.1. Simple horizontal scratches and trails
Carbonate O and C stable isotope compositions were analysed using
a Thermo Scientific Delta V Advance continuous flow isotope ratio mass 5.1.1. Ichnogenus Monomorphichnus Crimes, 1970 (Fig. 4 A–C)
spectrometer and a GasBench II preparation line connected to a Delta V This ichnogenus consists of horizontal straight to slightly sigmoidal
Advantage IRMS (Thermo Fisher Scientific) at the University of Tartu. subparallel ridges, preserved in positive epirelief. The ridges are ar­
Samples were chosen from the lithologically homogenous zones, and ca. ranged in pairs, with one ridge deeper than the other, and continual
2 mg of sample powder was dissolved in orthophosphoric acid overnight series that meet in a V. This trace fossil is unbranched, but crossovers or
at 70 ◦ C. The results of carbonate analyses are expressed in per mil de­ interpenetrations are found. The ridges are 10–13 cm long and are
viation relative to the Vienna PeeDee Belemnite (VPDB) scale for oxygen separated from each other by distances of 3–5 mm. The transverse ridges
(δ18O) and carbon (δ13C). Long-term reproducibility was better than ± are 2.0–3.0 mm wide. Central ridges are usually deeper than outer
0.2 ‰ (2σ) for δ18O and δ13C values. ridges.
These traces are tentatively assigned to Monomorphichnus Crimes,
1970. The presence of two main scratch marks at first sight makes this

4
A. Azizi et al. Precambrian Research 399 (2023) 107227

Fig. 4. A, subsurface of bed with arthropod scratch marks of the Monomorphichnus type (sample AA–MO–1). B & C, Magnified view of the boxed area in (A) show the
arrangement of parallel ridges and alternation of high and deeper ridges (arrowed). Scale bars are 1 cm.

form comparable to Monomorphichnus multilineatus, in which the central ichnogenus is preserved as concave epirelief along bedding planes in
ridges are deeper than the outer ridges (Gibb et al., 2017). Mono­ laminated siltstone.
morphichnus can be distinguished from Dimorphichnus by the lack of Gordia differs from Helminthopsis by its looped form (Pickerill and
pusher marks (Crimes, 1970; Osgood, 1970; Crimes et al., 1977). Jensen Peel, 1990). The Tabia examples exhibit increased meandering and
(1997) considered that both ichnogenera might represent different be­ crossing. Gordia is reported in various terrestrial and marine deposi­
haviours of the same producer. Monomorphichnus has been widely tional environments from several localities ranging in age from the late
documented in the Fortunian and younger strata (Mángano and Buatois, Ediacaran to Holocene (Mángano and Buatois, 2014; Turk et al., 2022).
2014, Marusin et al., 2021). Monomorphichnus-type scratch marks are Polychaetes or worm-like organisms have been proposed as possible
usually attributed to a trilobite tracemaker (Alpert, 1976; Sharma et al., tracemakers (Książkiewicz, 1977). Peel (2010) reported examples of
2018). ‘Gordia’-type traces on large arthropod shields from the early Cambrian
Sirius Passet fossil lagerstaette.
5.1.2. Ichnogenus Gordia Emmons, 1844 (Fig. 5 A-D)
The ichnogenus Gordia appears as a smooth meandering, un­ 5.1.3. Ichnogenus Helminthoidichnites Fitch, 1850 (Fig. 5C, D)
branched, horizontal trace, is uniformly 1.0–1.4 mm wide, and it The ichnogenus Helminthoidichnites appears as shallow, curved to
commonly has overcrossings, generating repetitive, looping spirals. This straight, 1.3–2.5 mm wide, unbranching trails, preserved as negative

5
A. Azizi et al. Precambrian Research 399 (2023) 107227

Fig. 5. Photographs and schematic tracing of the Tabia Member ichnofossils. A & B, Photograph and schematic tracing of Gordia isp. C & D, Photograph and
schematic tracing showing Helminthoidichnites isp and Gordia isp. (field photograph). Scale bars are 1 cm.

relief on the top surface of medium- to coarse-grained sandstone bed. 5.2. Simple actively filled (massive) horizontal to oblique structures
The trajectory of the trace is irregular. Overcrossing between several
burrows is common, and random loops are rare. The width is constant in 5.2.1. Ichnogenus Palaeophycus Hall, 1847 (Fig. 6A, B)
individual specimens. This ichnogenus has horizontal, smooth, cylindrical, straight to
Helminthoidichnites is one of the most common trace fossils from the slightly curved unbranched burrows that are 3–5 mm wide and up to
basal siltstone of the Tabia Member, and it is usually found in association 150 mm long, with a circular cross section. The walls are smooth and
with Gordia. Häntzschel (1975) regarded Helminthoidichnites as a junior nonornamented. The filling is the same as the host rock.
synonym of Gordia. However, other later studies (Buatois et al., 1998) Palaeophycus is commonly interpreted as a dwelling burrow of ver­
retained this ichnogenus. Helminthoidichnites differs from Gordia by miform animal deposit feeders, mostly polychaetes, usually moving
lacking self-overcrossing and from Helminthopsis in possessing a less parallel to the sediment surface (e.g., Pemberton and Frey, 1982; Hof­
winding nature (Hofmann and Patel, 1989). Helminthoidichnites range in mann et al., 2012; Knaust, 2017). Palaeophycus burrows can be distin­
age from the Ediacaran to the Holocene (e.g., Narbonne and Aitken, guished from Planolites burrows by the lined trace walls, as well as the
1990; Uchman et al., 2009; Turk et al., 2022). Vermiform animals are nature of the sediment filling. In contrast to Planolites, the filling of
suggested as potential trace makers of Helminthoidichnites-type traces Palaeophycus burrows is usually similar to the surrounding rock matrix
(Buatois et al., 1998; Evans et al., 2020). (Pemberton and Frey, 1982). Palaeophycus is an unbranched burrow;
however, pseudobranching is usually caused by the crossing of different

6
A. Azizi et al. Precambrian Research 399 (2023) 107227

Fig. 6. Photographs of the Tabia Member ichnofossils. A & B, Palaeophycus isp. (field photographs) The coin diameter is 2.4 cm. C, Planolites isp (sample; AA–Pl–1).
D, vertical cross section of Planolites isp. Scale bars are 1 cm.

burrow generations, which may result in misinterpretations and confu­ segments (Fig. 7F). Probes are segments that vary in length from 7.0 to
sion with other similar traces, such as Thalassinoides and Planolites 13.0 mm and in width from 2.1 to 3.0 mm. The traces change direction
(Weber et al., 2013). Palaeophycus-type traces have been reported from at the segment junctures, but no projections are observed. Segment
the late Ediacaran to Holocene (Scott et al., 2009; O’Neil et al., 2020). junctures occur as small depressions (Fig. 7C), although rare continued
and curved junctions are observed (Fig. 7F). Angles between segments
5.2.2. Ichnogenus Planolites Nicholson, 1873 (Fig. 6C-D) range from 95◦ to 165◦ .
The ichnogenus Planolites appears as horizontal to slightly oblique, Treptichnid traces have been previously reported from terminal
straight or irregularly sinuous, unbranched simple flattened cylinders, Ediacaran strata, such as in Spain (Jensen et al., 2007), Newfoundland
without ornamentation. It is elliptical in cross-section and 4–10 mm in (Gehling et al., 2001), Finnmark (Högström et al., 2013; McIlroy and
diameter, and the maximum preserved length is 10 cm. The burrow is Brasier, 2017), and Namibia (Darroch et al., 2016; Jensen and Runne­
filled with a material that is different from the host rock matrix; the gar, 2005; Jensen et al., 2000). Short-sized segments scattered on the
filling is finer-grained, dominated by quartz grains, while the sur­ bedding plane (Fig. 7A) share a similar morphology with those reported
rounding rock matrix is composed of quartz, feldspar, and mica flakes. from the Huns Member in Namibia but with more widely spaced ele­
Planolites is an actively filled burrow, interpreted as a feeding trace ments (Jensen et al., 2000; Darroch et al., 2021; Nelson et al., 2022), and
of vermiform deposit feeders (Pemberton and Frey, 1982, Rodríguez- those from the Cijara Formation in Spain (Jensen et al., 2007). However,
Tovar and Uchman, 2004; Pervesler et al., 2011). Worms, molluscs and uniserial probes arranged in a zigzag pattern are similar to those
arthropods are proposed as possible tracemakers of Planolites (Bromley, described by Jensen et al. (2007) and assigned to Treptichnid type
1996; Buatois and Mángano, 2002). Planolites have been described in traces. This complex structure may possibly represent three-dimensional
various continental and marine environments (e.g., Pemberton and burrow systems approaching the morphology of Treptichnus but is of
Frey, 1982; Rodríguez-Tovar and Uchman, 2004; Sarkar et al., 2009; lesser complexity. The relationship between Treptichnus and treptichnids
Phillips et al., 2011) and range in age from the late Ediacaran to Holo­ is uncertain (Jensen et al., 2007). It is unclear whether both traces have
cene (Häntzschel, 1975; Crimes, 1992; O’Neil et al., 2020). Recent in­ the same origin but are preserved differently or if treptichnids are
vestigations have revealed an absence of Planolites in the late Ediacaran similar to Treptichnus but represent less complex trace fossils. (see Jen­
(Mángano and Buatois, 2016; Buatois et al. 2020). sen et al., 2000). These occurrences provide insight into the potential
evolutionary origins of complex trace fossil behaviour and can be easily
5.3. Horizontal burrows with horizontal to vertical branches; Treptichnids incorporated into the criteria used in recognizing and correlating the
(Fig. 7A–F) Ediacaran–Cambrian boundary (Buatois et al., 2013; Landing et al.,
2013). The Tabia Member treptichnids represent the earliest form of
Treptichnid traces are common in the upper part of the Tabia ’complex’ trace fossil behaviour, albeit not as complex as that reflected
Member (3 m below the base of the Tifnout Member), with two distinct by T. pedum, the latter marking the base of the Cambrian (Brasier et al.,
morphologies observed. The first trace morphology (Fig. 7A, B) consists 1994). It is uncertain whether the simpler treptichnid burrows were
of individual probes that are generally oval in shape, with a maximum created by the same organisms as T. pedum, which is commonly thought
diameter of 6 mm. The second morphology (Fig. 7C–F) is composed of a to have been formed by priapulid worms (Vannier et al., 2010; Kesidis
subhorizontal series of individual, straight spindle-shaped probes ar­ et al., 2019).
ranged in a zigzag pattern. The shorter series consists of two segments
(Fig. 7E), whereas the longer series is composed of approximately five

7
A. Azizi et al. Precambrian Research 399 (2023) 107227

Fig. 7. Photographs of Treptichnids traces from the Tabia Member, all in sole view. A, Treptichnids on the sole of a sandstone bed showing individual probes oval in
shape (white arrows). B, Treptichnids on the sole of a sandstone bed showing individual probes of irregular shape (yellow arrows). C and D, Possible Treptichnids
composed of serial spindle-shaped probes arranged in a zigzagging pattern (white arrows); segment junctures occur as small depressions (yellow arrows). E and F,
Possible Treptichnids composed of serial straight probes arranged in a zigzagging pattern (white arrows); some segments are connected and show curved junctions
(yellow arrows). Scale bars are 1 cm.

5.4. Vertical plug-shaped burrows oriented perpendicular to the bedding plane. Fillings may reveal
patterned internal structures, such as chevron laminae, but not radial
5.4.1. Ichnogenus Conichnus Männil, 1966 (Fig. 8A, B and C) medusoid symmetry. Conichnus-type traces can be distinguished from
This ichnogenus has short, conical, vertical to slightly inclined bur­ Amphorichnus by the lack of amphora-like shape and the papillate
rows that are round to elliptical in transverse section. The studied termination. Amphorichnus also differs from Conichnus by the presence of
specimens show variable dimensions; the maximum diameter of the lateral adjustment traces (Vinn et al., 2015). Conichnus traces have been
burrows is 23 mm, and they are up to 25 mm deep, with a resulting recorded from the lower Cambrian to the Cenozoic (Pemberton et al.,
diameter/depth ratio of 0.92. The burrow fill is composed of fine- 1988). Plug-shaped burrows interpreted as Conichnus are the most
grained siliciclastic–carbonate sediments. The lithological difference common trace fossils found in the late Ediacaran strata of Namibia
between the burrow fill and the host rock is clearly marked. The central (Darroch et al., 2016; Cribb et al., 2019; Darroch et al., 2021). Con­
cone is rich in siliciclastic material dominated by fine-grained quartz ichnus-type traces are interpreted as the dwelling burrows of a soft-
grains (Fig. 8C), which is somewhat more resistant to weathering than bodied anemone-like organism (Frey and Howard, 1985, Pemberton
the surrounding part and the host sediments. Concentric laminae are et al., 1988). The variously filled internal structures of Conichnus bur­
observed on the top (Fig. 8B). The basal part may be slightly rounded or rows suggest that the trace maker moved periodically upwards as the
conical. animal tried to keep pace with sedimentation (Frey and Howard, 1985).
In certain aspects, the majority of these conical traces resemble to the
ichnogenera Conichnus Männil, (1966). According to Männil (1966), the 5.4.2. Problematic paired plug-shaped burrows (Fig. 9 A, B and C)
ichnogenus Conichnus comprises acuminated subcylindrical structures Pairs of plug-shaped holes are very common at the boundary

8
A. Azizi et al. Precambrian Research 399 (2023) 107227

Fig. 8. Trace fossil specimens of the Conichnus type from the boundary between the Tabia and Tifnout Members are shown. A, a field photograph of Conostichus in
sandy carbonate mixed sediments (sample: AA–Co–1). B, a top view of Conostichus showing concentric laminae (arrowed) (sample: AA–Co–2). C, a vertical view of
Conichnus displaying a conical shape. Scale bars are 1 cm.

between the Tabia and Tifnout members and may occur in association Zaouia section includes abundant horizontal burrows and trails, prob­
with most of the other traces described above. They are usually pre­ ably produced by priapulids (Treptichnids) and arthropods (Mono­
served on bedding planes as a pair of holes and look superficially like the morphichnus), as well as by anemone-like organisms (Conichnus). The
Arenicolites-type trace, but a “U” shape has not been detected in cross- ichnotaxonomic composition and stratigraphic distribution of the trace
section, and this type of trace does not exceed 4 mm in depth. The di­ fossils of this part of the Adoudou Formation are consistent with For­
ameters of the holes are 4–12 mm, and the spacing between the holes of tunian ichnoassemblages (Mángano and Buatois, 2017).
each pair varies between 4 and 10 mm. The Ediacaran–Cambrian boundary is marked by the first appear­
Paired plug-shaped burrows have been previously described within ance datum (FAD) of T. pedum, the first three-dimensionally branching
the E-C boundary interval in the Nama Group, Namibia (Darroch et al., burrows (Narbonne et al., 1987; Brasier et al., 1994; Landing, 1994;
2016; Cribb et al., 2019; Darroch et al., 2021). As mentioned by Darroch Buatois et al., 2013), as well as other biostratigraphic ichnofossils that
et al. (2021), the neighbouring holes seem to be alternating or aligned, made their first appearance at or immediately above the boundary, such
suggesting affinities with treptichnids. Similar plug-shaped forms are as Arenicolites sp., Conichus conicus, Monomorphichnus spp., Phycodes sp.,
interpreted as Brooksella by Crimes and Germs (1982). However, the or Helminthopsis tenius (Jensen, 2003; Landing et al., 2013; Babcock
association of burrows in pairs remains mysterious. et al., 2014).
The FAD of T. pedum roughly corresponds to the base of the
6. Discussion Cambrian worldwide, including the Mackenzie Mountains of Canada
(Carbone and Narbonne, 2014); Sonora, Mexico (Sour-Tovar et al.,
The trace fossil record provides a key piece of evidence to evaluate 2007); Death Valley, Eastern California (Jensen et al., 2002); Flinders
the timing of diversification and ecosystem construction during the Ranges of South Australia (Jensen et al., 1998); eastern Finnmark,
terminal Ediacaran–early Cambrian. Ediacaran shallow marine deposits Norway (Føyn and Glaessner, 1979); and Nama Group, Namibia (Lin­
worldwide contain simple horizontal trace fossils reflecting the activity nemann et al., 2019; Darroch et al., 2021), among many other regions.
of nonspecialized microbial mat grazers (Buatois and Mángano, 2012; Nevertheless, the selection of T. pedum as a proxy for the base of the
Darroch et al., 2021). The Cambrian radiation of complex animal life Cambrian period has also been heavily criticized (Babcock et al., 2014;
induced a dramatic increase in the depth and complexity of bioturbation Zhu et al., 2019). On the one hand, the three-dimensional structure of
in the seafloor sediment, known as the ‘agronomic revolution’ (AR). The T. pedum remains the subject of debate (Geyer, 2005; Seilacher, 2007;
AR marks the shift from matgrounds to bioturbated mixgrounds (Buatois Vannier et al., 2010; Chen et al., 2013; Meyer et al., 2014). According to
et al., 2014). The timing of this geobiological event is still a matter of Seilacher (2007), the trace fossil T. pedum is indicative of a sophisticated
debate, whereas significant diversification of trace fossils and the advent trace of undermat mining rather than the vertical bioturbation proposed
of bioturbation occurred ca. 20 Myr earlier than the “Cambrian Explo­ by Geyer and Uchman (1995) and Geyer (2005). Undermat mining
sion” of body fossils (Buatois and Mangano, 2016). forms are reported in the Ediacaran Shibantan Member of the Dengying
The transition between the Tabia and the Tifnout members in the Formation in South China (Meyer et al., 2014), as well as in the lower

9
A. Azizi et al. Precambrian Research 399 (2023) 107227

Fig. 9. Trace fossil specimens of the paired vertical plug-shaped burrows from the boundary between the Tabia and Tifnout Members. A and B, slab specimens
showing numerous pairs of holes (sample; AA–PH–1). B & C, vertical cross–section of the holes showing in (A), the position of holes is marked by white arrows. Scale
bars are 1 cm.

Member of the Wood Canyon Formation in the Great Basin, United Ediacaran–Cambrian sections (Buatois et al., 2013; Shahkarami et al.,
States (O’Neil et al., 2020). In addition, the morphology of Treptichnids is 2017; Buatois, 2018; Devaere et al., 2021). The Fortune Head stratotype
not yet well understood, making it difficult to make a definitive taxo­ is a coarse siliciclastic-dominated sequence that has undergone meta­
nomic determination of them in relation to the ichnogenus Treptichnus. morphism and lacks a chemostratigraphic, magnetostratigraphic, or
Jensen et al. (2000) described Treptichnids in more detail in the base of other non-biostratigraphic record that would enable correlation with
the Huns Member of the Nama Group and identified them as Treptichnus other sections regionally or globally (Geyer, 2005). The FAD of T. pedum
isp. is inferred to lie above the interval that contains mostly Ediacaran fossils
On the other hand, the facies dependence of T. pedum has been and below the interval containing mostly Cambrian fossils. This section
regarded as a major complication in its use for global correlation has not been proven to be the stratotype; however, it is strongly sup­
(Gehling et al., 2001; Geyer, 2005; Babcock et al., 2014; Zhu et al., ported by numerous studies around the world (Geyer, 2005; Maloof
2019). Generally, these trace fossils are found predominantly in silici­ et al., 2005, 2010; Landing et al., 2013). It should also be considered that
clastic rocks, and although they are known to occur in a range of lith­ Ediacaran-type macrofossils have rarely been reported from lower
ofacies (Landing et al., 2013; Buatois et al., 2013; Buatois, 2018), Cambrian strata. For example, Ediacaran-type fronds were reported
carbonate facies are usually not included. This limitation reduces the from the Uratanna Formation in the Angepena syncline, northern Flin­
application of T. pedum as a proxy for the base of the Cambrian, as this ders Ranges, above the Ediacaran–Cambrian transition (Jensen et al.,
interval often consists of carbonate facies in numerous key regions, such 1998). Simple discoidal fossils and frond-like soft-bodied fossils are
as Kazakhstan, China, Mongolia, and Siberia. Facies dependence may described from the Cambrian strata of the Salt Spring Hills and the White
explain the delayed appearance of T. pedum in some Mountains (Hagadorn et al., 2000). In addition, Cloudinomorphs and

10
A. Azizi et al. Precambrian Research 399 (2023) 107227

Table 1 Ediacaran–Cambrian transition to be somewhat speculative in practice.


Carbon and oxygen isotope data. Nonetheless, the first appearance of typical Fortunian trace fossils allows
Lab. ID ID1 ID2 δ13C(‰V-PDB) δ18O(‰V-PDB) us to constrain the Cambrian age for the base of the Tifnout Member.
This is also supported by the carbon isotopic δ13Ccarb trend in the studied
0948-22 Ti 1 − 3,75 9,40
succession, showing δ13Ccarb values in the basal limestone unit aver­

0949-22 Ti 2 − 3,82 − 9,38
0950-22 Ti 3 − 3,70 − 9,50 aging 0 ‰ but staying rather stable at − 3 to − 4 ‰ in carbonates of the
0951-22 Ti 4 − 3,50 − 9,33 Tifnout Member (Fig. 2). A large negative δ13C excursion, the BACE, has
0952-22 Ti 5 − 3,47 − 8,59 been recorded worldwide postdating the last appearance of Ediacaran-
0953-22 Ti 6 − 3,34 8,03
type fossils and/or predating the first occurrence of T. pedum (Zhang

0954-22 Ti 7 − 3,26 − 8,83
0955-22 Ti 8 − 3,20 − 9,19 et al., 1997; Corsetti and Hagadorn, 2000; Amthor et al. 2003; Smith
0956-22 Ti 9 − 3,07 − 9,04 et al., 2016; Chen and Feng, 2019). The BACE is known from Mongolia
0957-22 Ti 10 − 3,17 − 8,51 (Brasier et al., 1996; Smith et al., 2016), the Siberian platform (Kaufman
0958-22 Ti 11 − 3,38 9,00

and Knoll, 1995; Bartley et al., 1998; Kouchinsky et al., 2017), north­
0959-22 Ti 12 − 3,23 − 8,40
0960-22 Ti 13 − 3,05 − 9,60 west Canada (Narbonne et al., 1994), Iran (Kimura et al., 1997), the
0961-22 Ti 14 − 3,07 − 9,26 Great Basin (Corsetti and Hagadorn, 2000; Corsetti and Kaufman, 2003;
0962-22 Ti 15 − 3,20 − 9,44 Smith et al., 2016), Kazakhstan (Stammeier et al., 2019), South China
0963-22 Ti 16 − 2,97 − 10,22 (Ishikawa et al., 2008, 2014; Li et al., 2009; Steiner et al., 2020), Mexico
0964-22 Ti 17 − 3,00 10,39
(Loyd et al., 2012, 2013; Hodgin et al., 2021) and Morocco (Maloof

0965-22 Ti 18 − 3,22 − 10,29
0966-22 Ti 19 − 3,09 − 9,69 et al., 2005, 2010; and herein). Therefore, this negative excursion is
0967-22 Ti 20 − 2,95 − 10,98 thought to mark the Ediacaran–Cambrian transition and has been linked
0968-22 Ti 21 − 3,10 − 10,14 to the perturbation of the carbon cycle and, possibly, a mass extinction
0969-22 Ti 22 − 3,07 12,01

(Amthor et al., 2003; Darroch et al., 2018; Hodgin et al., 2021);
0970-22 Ti 23 − 2,78 − 10,51
0971-22 Ti 24 − 3,18 − 11,80 nevertheless, its absolute age has not been clearly defined. Geochrono­
0972-22 Ti 25 − 3,44 − 11,30 logical dating of the Ediacaran–Cambrian excursion comes from the Ara
0973-22 Cb 1 − 0,81 − 7,12 Group in Oman, where the U–Pb ages obtained from a volcanic ash bed
0974-22 Cb 2 − 1,01 − 8,25 that coincides with the negative excursion yield 542.0 ± 0.3 Ma
0975-22 Cb 3 − 0,78 6,85
(Amthor et al., 2003) and are recalibrated to 541.00 ± 0.13 Ma by

0976-22 Cb 4 − 1,67 − 5,34
0977-22 Cb 5 − 0,89 − 8,41 Bowring et al. (2007). Recent U-Pb ages obtained for a volcanic ash bed
0978-22 Cb 6 − 0,85 − 8,78 located three metres above the BACE in southern China have yielded
0979-22 Cb 7 − 0,39 − 6,85 540.7 ± 3.8 Ma (Chen and Feng, 2019), providing further support for
0980-22 Cb 8 0,29 6,59

the validity of the overall excursion of the Ediacaran–Cambrian
0981-22 Cb 9 − 0,08 − 7,16
0982-22 Cb 10 − 0,33 − 7,40 boundary.
0983-22 Cb 11 0,18 − 7,77 The composite carbon isotope curve of the Sous Basin shows two
0984-22 Cb 12 0,02 − 7,31 negative ‘W’-shaped excursions at the base of the Adoudou Formation
0985-22 Cb 13 − 0,05 − 7,53 (e.g., Maloof et al., 2005, 2010), where the lower anomaly (~− 4 ‰) is
0986-22 Cb 14 − 0,25 7,43
recorded in the Basal limestone unit of the Tabia Member, and the upper

0987-22 Cb 15 0,93 − 7,55
0988-22 Cb 16 1,32 − 8,47 excursion (~− 6 ‰) oscillates in the lower fifty metres of the Tifnout
0989-22 Cb 17 − 0,97 − 6,64 Member. Identical ‘W’-shaped δ13Ccarb signatures have been recorded in
0990-22 Cb 18 0,32 − 8,64 the Siberian Platform (Brasier et al. 1996) and SW Mongolia (Smith
0991-22 Cb 19 − 0,25 6,63

et al., 2016). In the Zaouia area, this upper excursion is recorded in the
0992-22 Cb 20 − 0,23 − 8,20
0993-22 Cb 21 − 0,39 − 8,07
base of the Tifnout Member and regarded as marking the Ediacar­
0994-22 Cb 22 − 2,61 − 6,50 an–Cambrian transition of the Sous Basin (e.g., Maloof et al., 2005,
0995-22 Cb 23 2,22 − 9,37 2010). The carbonate isotope values in the studied profile suggest,
0996-22 Cb 24 1,62 − 9,94 however, that basal limestone with nearly normal marine values is
chemostratigraphically below the ‘W’-shaped excursion, and the sam­
Erniettomorphs are found in the Nomtsas Formation in Namibia, and they ples from the Tifnout Member carbonates overlying the basal siltite unit
are dated to 538.56 ± 0.09 Ma and stratigraphically overlap complex with the observed ichnofossils show values and a rather flat stratigraphic
bilaterian trace fossils, such as Archaeonassa, Psammichnites, Para­ trend corresponding to the lower δ13Ccarb anomaly at the beginning of
psammichnites, and treptichnids (Nelson et al., 2022). Moreover, dis­ the ‘W’-shaped excursion. Therefore, based on the C isotope character­
coidal fossils were previously described from the Tabia Member in the istics and ichnofossil occurrence, the Ediacaran–Cambrian boundary can
Warmandaz region, which neighbours the Igherm inlier and is located be placed at the base of the Tifnout Member and below the most nega­
approximately 30 km eastwards from the studied section (Houzay, 1979; tive part of the BACE. This conclusion is further justified by the last
Maloof et al., 2005, 2010; Letsch et al., 2019). These are 1–13 cm wide occurrences of Treptichnids and Ediacaran-type fossils in the Tabia
and preserved both as positive epireliefs on the tops of beds and corre­ Member.
sponding positive hyporelief casts on the soles of beds. Following the In addition, the Ediacaran–Cambrian transition coincides with a
argumentation of Letsch et al. (2019), the Warmandaz specimens share major unconformity observed in many sections: Mackenzie Mts., Canada
numerous characteristics with the genus Aspidella, which is described in (Narbonne et al., 1994); Death Valley, United States of America (Corsetti
many Ediacaran strata around the world (Tarhan et al., 2015). Never­ and Kaufman, 1994; Corsetti and Hagadorn, 2000); and the Ara Group,
theless, it is evident that the decline of Ediacaran fauna is coincident Oman (Amthor et al., 2003). Additionally, this major discontinuity is
with the rise of bioturbation, also hinting that sediment disturbance may located below and not far from the BACE in Lena-Aldan, Siberia (Brasier
have played a prominent role in the decline of the soft-bodied Ediacaran et al., 1994; Pelechaty et al., 1996), and SW Mongolia (Brasier et al.,
organisms (Mángano and Buatois, 2014; Darroch et al., 2015). 1996; Smith et al., 2016). Similarly, a major discontinuity in the Sous
Therefore, the facies dependence of T. pedum, as well as the strati­ Basin is located at the base of the Tabia Member (Fig. 1C), which is in
graphic overlaps of complex bilaterian trace fossils with Ediacaran body agreement with the C isotope and fossil data.
fossils cause both the localization and the correlation of the

11
A. Azizi et al. Precambrian Research 399 (2023) 107227

7. Conclusion Baidada, B., Ikenne, M., Barbey, B., Soulaimani, A., Cousens, B., Haissen, F., Ilmen, S.,
Alansari, A., 2018. SHRIMP U-Pb zircon geochronology of the granitoids of the
Imiter Inlier: Constraints on the Pan-African events in the Saghro massif, Anti-Atlas
The Ediacaran–Cambrian boundary has long been recognized as an (Morocco). J. Afr. Earth Sc. 150, 799–810.
important junction in Earth history. Despite this, it is still a poorly un­ Baidder, L., Michard, A., Soulaimani, A., Fekkak, A., Eddebbi, A., Rjimati, E.C., Raddi, Y.,
derstood geological transition due to limited geochronological data and 2016. Fold interference pattern in thick-skinned tectonics; a case study from the
external Variscan belt of Eastern Anti-Atlas, Morocco. J. Afr. Earth Sci. 119,
the difficulty in correlating stratigraphic sections regionally and 204–225.
globally. Banner, J.L., Hanson, G.N., 1990. Calculation of simultaneous isotopic and trace element
The first occurrence of the Cambrian trace fossils from the base of the variations during water-rock interaction with applications to carbonate diagenesis.
Geochimica et Cosmochimica Acta, 54, 3123–3137.
Tifnout Member allows us to suggest at least a Fortunian age for the host Bartley, J.K., Pope, M., Knoll, A.H., Semikhatov, M.A., Petrov, P.Y., 1998. A Vendian
strata. Further evidence of Treptichnid traces found 3 m below the base Cambrian boundary succession from the northwestern margin of the Siberian
of the Tifnout Member supports a late Ediacaran age for the Tabia Platform: stratigraphy, palaeontology, chemostratigraphy and correlation. Geol.
Mag. 135, 473–494.
Member. Benssaou, M., Hamoumi, N., 2001. L’Anti-Atlas occidental du Maroc: étude
The integration of trace fossils with carbon isotope δ13C records sédimentologique et reconstitutions paléogéographiques au Cambrien inférieur.
suggests that the Ediacaran–Cambrian transition of the Sous Basin may J. Afr. Earth Sc. 32, 351–372.
Benssaou, M., Hamoumi, N., 2004. Les microbialites de l’Anti-Atlas occidental (Maroc):
have occurred at the contact between the Tifnout and Tabia Members, marqueurs stratigraphiques et témoins des changements environnementaux au
chemostratigraphically corresponding to the large negative δ13Ccarb Cambrien inférieur. C. R. Geosci. 336, 109–116.
excursion that was previously suggested for the Ediacaran–Cambrian Blein, O., Baudin, T., Chévremont, P., Soulaimani, A., Admou, H., Gasquet, D.,
Cocherie, A., Egal, E., Youbi, N., Razin, P.h., Bouabdelli, M., Gombert, P.h., 2014.
boundary (BACE) in the western Anti-Atlas Mountains.
Geochronological constraints on the polycyclic magmatism in the Bou Azzer-El
Graara inlier (central Anti-Atlas Morocco). J. Afr. Earth Sc. 99, 287–306.
CRediT authorship contribution statement Bottjer, D.J., Hagadorn, J.W., Dornbos, S.Q., 2000. The Cambrian substrate revolution.
GSA Today 10, 1–7.
Bowring, S.A., Grotzinger, J.P., Condon, D.J., Ramezani, J., Newall, M.J., Allen, P.A.,
Abdelfattah Azizi: Conceptualization, Formal analysis, Funding 2007. Geochronologic constraints on the chronostratigraphic framework of the
acquisition, Investigation, Project administration, Writing – original Neoproterozoic Huqf Supergroup, Sultanate of Oman. Am. J. Sci. 307, 1097–1145.
Brasier, M., Cowie, J., Taylor, M., 1994. Decision on the Precambrian-Cambrian
draft, Writing – review & editing. Asmaa El Bakhouch: . Abderrazak El
boundary stratotype. Episodes 17, 3–8.
Albani: Writing – review & editing. Kalle Kirsimäe: Writing – review & Brasier, M., Dorjnamjaa, D., Lindsay, J., 1996. The Neoproterozoic to early Cambrian in
editing. Mouhssin El Halim: . Khadija El Hariri: Writing – review & southwest Mongolia: an introduction. Geol. Mag. 133, 365–369.
Bromley, R.G., 1996. Trace Fossils. Biology, Taphonomy and Applications. Chapman and
editing. Mohamed Erragragui: . Ahmid Hafid: . Olev Vinn: Writing –
Hall, London, p. 361.
review & editing. Buatois, L.A., 2018. Treptichnus pedum and the Ediacaran-Cambrian boundary:
significance and caveats. Geol. Mag. 155, 174–180.
Buatois, L.A., Mángano, M.G., Maples, C.G., Lanier, W.P., 1998. Allostratigraphic and
Declaration of competing interest sedimentologic applications of trace fossils to the study of incised estuarine valleys:
an example from the Virgilian Tonganoxie Sandstone member of eastern Kansas.
Current Res. Earth. Sci. 241, 1–27.
The authors declare that they have no known competing financial Buatois, L.A., Almond, J., Germs, G.J.B., 2013. Environmental tolerance and range offset
interests or personal relationships that could have appeared to influence of Treptichnus pedum: implications for the recognition of the Ediacaran-Cambrian
the work reported in this paper. boundary. Geology 41, 519–522.
Buatois, L.A., Mángano, M.G., 2012. The trace-fossil record of organism-matground
interactions in space and time. In: Noffke, N., Chafetz, H. (Eds.), Microbial Mats in
Data availability Siliciclastic Depositional Systems through Time. Society for Sedimentary Geology,
Special Publication. 101, pp. 15–28.
Buatois, L.A., Mangano, M.G., 2016. Ediacaran ecosystems and the dawn of animals. In:
No data was used for the research described in the article. M angano, M.G., Buatois, L.A. (Eds.), The Trace-fossil Record of Major Evolutionary
Events, Topics in Geobiology, 39. doi: 10.1007/978-94-017-9600-2_2.
Acknowledgments Buatois, L.A., Mángano, M.G., Minter, N.J., Zhou, K., Wisshak, M., Wilson, M.A., Olea, R.
A., 2020. Quantifying ecospace utilization and ecosystem engineering during the
early Phanerozoic – the role of bioturbation and bioerosion. Sci. Adv. 6 eabb0618.
The authors are grateful to anonymous reviewers for their doi: 10.1126/sciadv.abb0618.
constructive comments. We thank Cadi Ayyad University, Académie Buatois, L.A., Mángano, M.G., 2002. Trace fossils from Carboniferous floodplain deposits
in western Argentina: implications for ichnofacies models of continental
Hassan II des Sciences et Techniques and the Region Nouvelle Aquitaine environments. Palaeogeogr. Palaeoclimatol. Palaeoecol. 183, 71–86.
for their financial support. The authors are grateful to Gabriela Buatois, L.A., Narbonne, G.M., Mángano, M.G., Carmona, N.B., Myrow, P., 2014.
Mángano, Luis Buatois, E.H. Bouougri and Ursula Toom for their help in Ediacaran matground ecology persisted into the earliest Cambrian. Nat. Commun. 5,
1–5.
trace fossils identification. Idir Elhabib, and Ibrahim Ouargaga are Carbone, C., Narbonne, G.M., 2014. When life got smart: the evolution of behavioral
thanked for their assistance during the field work. The analyses at the complexity through the Ediacaran and early Cambrian of NW Canada. J. Paleo. 88,
University of Tartu were supported by the Estonian Centre of Analytical 309–330.
Chen, Z., Zhou, C., Meyer, M., Xiang, K., Schiffbauer, J.D., Yuan, X., Xiao, S., 2013. Trace
Chemistry.
fossil evidence for Ediacaran bilaterian animals with complex behaviors. Precambr.
Res. 224, 690–701.
References Chen, C., Feng, Q., 2019. Carbonate carbon isotope chemostratigraphy and U-Pb zircon
geochronology of the Liuchapo Formation in South China: constraints on the
Ediacaran-Cambrian boundary in deep-water sequences. Palaeogeogr.
Alpert, S.P., 1976. Planolites and Skolithos from the Upper Precambrian-Lower Cambrian
Palaeoclimatol. Palaeoecol. 535, 109361.
White-Inyo mountains, California. J. Paleontol. 49, 508–521.
Choubert, G., 1947. L’accident majeur de l’Anti-Atlas. Comptes Rendus Hebdomadaires
Álvaro, J.J., Clausen, S., 2006. Microbial crusts as indicators of stratigraphic diastems in
des Seances de l’Académie des Sciences. 224, 1172–1173.
the Cambrian Micmacca Breccia, Moroccan Atlas. Sed. Geol. 185, 255–265.
Choubert, G., 1963. Histoire géologique du Précambrien de l’Anti-Atlas. Not. Mém. Serv.
Álvaro, J.J., Debrenne, F., 2010. The Great Atlasian Reef Complex: an early Cambrian
Géol. Maroc. 162, 1–352.
subtropical fringing belt that bordered West Gondwana. Palaeogeogr.
Choubert, G., 1952. Histoire géologique du domaine de l’Anti-Atlas. In: Choubert, G.,
Palaeoclimatol. Palaeoecol. 294, 120–132.
Marçais, J. (Eds.), Géologie du Maroc. 19th Int. Geol. Congr., Alger 1952. Monogr.
Amthor, J., Grotzinger, J., Schröder, S., Bowring, S., Ramezani, J., Martin, M., Matter, A.,
Rég. Sér. 3 (Maroc). 6, 77–194.
2003. Extinction of Cloudina and Namacalathus at the Precambrian-Cambrian
Clauer, N., 1974. Utilisation de la methode rubidium-strontium pour la datation d’une
boundary in Oman. Geology 31, 431–434.
schistosite de sédiments peu métamorphisés: application au Precambrien II de la
Azizi, A., El Albani, A., El Bakhouche, A., Vinn, O., Bankole, O.M., Fontaine, C., Hafid, A.,
boutonniere de BouAzzer-El Graara (Anti-Atlas, Maroc). Earth Planet. Sci. Lett. 22,
Kouraiss, K., El Hariri, K., 2022. Early biomineralization and exceptional
404–412.
preservation of the first thrombolite reefs with archaeocyaths in the lower Cambrian
Clausen, S., Álvaro, J.J., Zamora, S., 2014. Replacement of benthic communities in two
of the western Anti-Atlas, Morocco. Geol. Mag. 1–16.
Neoproterozoic-Cambrian subtropical-to-temperate rift basins, High Atlas and Anti-
Babcock, L.E., Peng, S., Zhu, M., Xiao, S., Ahlberg, P., 2014. Proposed reassessment of the
Atlas, Morocco. J. Afr. Earth Sci. 98, 72–93.
Cambrian GSSP. J. Afr. Earth Sc. 98, 3–10.

12
A. Azizi et al. Precambrian Research 399 (2023) 107227

Corsetti, F.A., Hagadorn, J.W., 2000. Precambrian-Cambrian transition: Death Valley, Häntzschel, W., 1975. Treatise on invertebrate paleontology / Part W. Miscellanea.
United States. Geology 28, 299–302. Supplement 1, Trace fossils and problematica. Geological Society of America.
Corsetti, F.A., Kaufman, A.J., 1994. Chemostratigraphy of Neoproterozoic-Cambrian (University of Kansas Press, Boulder, Colorado, Lawrence). Pp. 269.
units, White-Inyo region, eastern California and western Nevada: Implications for Hefferan, K., Soulaimani, A., Samson, S., Admou, H., Inglis, J., Saquaque, A., Latifa, C.,
global correlation and faunal distribution. PALAIOS 211–219. Heywood, N., 2014. A reconsideration of Pan African orogenic cycle in the Anti-Atlas
Corsetti, F.A., Kaufman, A.J., 2003. Stratigraphic investigations of carbon isotope Mountains, Morocco. J. Afr. Earth Sci. 98, 34–46.
anomalies and Neoproterozoic ice ages in Death Valley, California. Geol. Soc. Am. Hodgin, E.B., Nelson, L.L., Wall, C.J., Barron-Díaz, ́ A.J., Webb, L.C., Schmitz, M.D., Fike,
Bull. 115, 916–932. D.A., Hagadorn, J.W., Smith, E.F., 2021. A link between rift-related volcanism and
Cribb, A.T., Kenchington, C.G., Koester, B., Gibson, B.M., Boag, T.H., Racicot, R.A., end-Ediacaran extinction? Integrated chemostratigraphy, biostratigraphy, and U-Pb
Mocke, H., Laflamme, M., Darroch, S.A., 2019. Increase in metazoan ecosystem geochronology from Sonora, Mexico. Geology. 49, 115–119.
engineering prior to the Ediacaran-Cambrian boundary in the Nama Group, Hofmann, R., Mángano, M. G., Elicki, O., Shinaq, R., 2012. Paleoecologic and
Namibia. R. Soc. Open Sci. 6, 190548. Biostratigraphic Significance of Trace Fossils From Shallow- to Marginal-Marine
Crimes, T.P., 1970. The significance of trace fossils in sedimentology, stratigraphy and Environments From the Middle Cambrian (Stage 5) of Jordan. Cambridge University
palaeoecology with examples from Lower Palaeozoic strata. Trace Fossils 2, Press (CUP). J. Paleontol. 86, 931–955.
101–126. Hofmann, H.J., Patel, I.M., 1989. Trace fossils from the type ‘Etcheminian Series’ (Lower
Crimes, T.P., 1992. Changes in the trace fossil biota across the Proterozoic-Phanerozoic Cambrian Ratcliffe Brook Formation), Saint John area, New Brunswick, Canada.
boundary. J. Geol. Soc. Lond. 149, 637–646. Geol. Mag. 126, 139–157.
Crimes, T.P., Legg, I., Marcos, A., Arboleya, M., 1977. Late Precambrian– low Lower Högström, A., Jensen, S., Palacios, T., Ebbestad, J.O.R., 2013. New information on the
Cambrian trace fossils from Spain. In: Crimes, T.P., Harper, J.C. (Eds.), Trace fossils Ediacaran-Cambrian transition in the Vestertana Group, Finnmark, northern
2: Geological Journal, Special Issue 9, 91–138. Norway, from trace fossils and organic-walled microfossils. Nor. J. Geol. 93, 95–106.
Crimes, T.P., Germs, G.J.B., 1982. Trace fossils from the Nama Group Houzay, J.P., 1979. Empreintes attribuables à des méduses dans la série de base de
(PrecambrianCambrian) of Southwest Africa (Namibia). J. Paleo. 56, 890–907. l’Adoudounien (Précambrien terminal de l’Anti-Atlas Maroc). Géologie
Darroch, S.A., Boag, T.H., Racicot, R.A., Tweedt, S., Mason, S.J., Erwin, D.H., Méditerranéenne. VI/3, 379–384.
Laflamme, M., 2016. A mixed Ediacaran-metazoan assemblage from the Zaris Sub- Hupé, P., 1960. Sur le Cambrien inférieur du Maroc. Rep. In: 21st Int. Geol. Congr.,
basin, Namibia. Palaeogeogr. Palaeoclimatol. Palaeoecol. 459, 198–208. Norden. 8, 75–85.
Darroch, S.A., Cribb, A.T., Buatois, L.A., Germs, G.J., Kenchington, C.G., Smith, E.F., Ikenne, M., Soderlund, U., Ernst, R, E., Pin, C., Youbi, N., El Aouli, E, H., Hafid, A., 2017.
Mocke, H., O’Neil, G.R., Schiffbauer, J.D., Maloney, K.M., Racicot, R.A., Turk, K.A., A c. 1710 Ma mafic sill emplaced into a quartzite and calcareous series from Ighrem,
Gibson, B.M., Almond, J., Koester, B., Boag, T.H., Tweedt, S.M., Laflamme, M., 2021. Anti-Atlas-Morocco: Evidence that the Taghdout passive margin sedimentary group
The trace fossil record of the Nama Group, Namibia: exploring the terminal is nearly 1 Ga older than previously thought. J. Afr. Earth Sci. 127, 62–76.
Ediacaran roots of the Cambrian explosion. Earth Sci. Rev. 212, 103435. Ishikawa, T., Ueno, Y., Shu, D., Li, Y., Han, J., Guo, J., Yoshida, N., Maruyama, S.,
Darroch, S.A.F., Smith, E.F., Laflamme, M., Erwin, D.H., 2018. Ediacaran extinction and Komiya, T., 2014. The δ13C excursions spanning the Cambrian explosion to the
Cambrian explosion. Trends Ecol. Evol. 33, 653–663. Canglangpuian mass extinction in the Three Gorges area, South China. Gondwana
Darroch, S.A., Sperling, E.A., Boag, T.H., Racicot, R.A., Mason, S.J., Morgan, A.S., Res. 25, 1045–1056.
Tweedt, S., Myrow, P., Johnston, D.T., Erwin, D.H., Laflamme, M., 2015. Biotic Ishikawa, T., Ueno, Y., Komiya, T., Sawaki, Y., Han, J., Shu, D., Li, Y., Maruyama, S.,
replacement and mass extinction of the Ediacara biota. Proc. R. Soc. B, 20151003. Yoshida, N., 2008. Carbon isotope chemostratigraphy of a Precambrian/Cambrian
Debrenne, F., Debrenne, M., 1995. Archaeocyaths of the Lower Cambrian of Morocco. boundary section in the Three Gorge area, South China: prominent global-scale
Beringeria. Spec. Iss. 2, 121–145. isotope excursions just before the Cambrian Explosion. Gondwana Res. 14, 193–208.
Devaere, L., Korn, D., Ghaderi, A., Struck, U., Bavandpur, A.K., 2021. New and revised Jensen, S., 1997. Trace fossil from the Lower Cambrian Mickwitzia sandstone, south-
Small Shelly Fossil record from the Lower Cambrian of Northern Iran. Papers. central Sweden. Fossils Strata 42, 1–110.
Palaeontology. 7, 2141–2181. Jensen, S., Droser, M.L., Heim, N.A., 2002. Trace fossils and ichnofabrics of the Lower
El Hadi, H., Simancas, F., Martìnez-Poyatos, J., Azor, A., Tahiri, A., Montero, P., Fanning, Cambrian Wood Canyon Formation, southwest Death Valley area. In: Coresetti, F.A.
C.M., Bea, F., Gonz_alez-Lodeiro, F., 2010. Structural andgeochronological (Ed.), Proterozoic-Cambrian of the Great Basin and Beyond. Pacific Section SEPM,
constraints on the evolution of the Bou Azzer Neoproterozoic ophiolite (Anti-Atlas, pp. 123–135.
Morocco). Precambr. Res. 182, 1–14. Jensen, S., Palacios, T., Martí Mus, M., 2007. A brief review of the fossil record of the
Emmons, E., 1844. The Taconic System; Based on Observations in New York, Ediacaran-Cambrian transition in the area of Montes de Toledo-Guadalupe, Spain.
Massachusetts, Maine, Vermont and Rhode Island. Carroll and Cook Printers, Geol. Soc. Lond. Spec. Publ. 286, 223–235.
Albany, New York, pp. 1–68. Jensen, S., Runnegar, B.N., 2005. A complex trace fossil from the Spitskop Member
Erwin, D.H., 1999. The origin of bodyplans. Am. Zool. 39, 617–629. (terminal Ediacaran–? Lower Cambrian) of southern Namibia. Geol. Mag. 142,
Evans, S.D., Hughes, I.V., Gehling, J.G., Droser, M.L., 2020. Discovery of the oldest 561–569.
bilaterian from the Ediacaran of South Australia. Proc. Natl. Acad. Sci. 117, Jensen, S., Gehling, J.G., Droser, M.L., 1998. Ediacara-type fossils in Cambrian
7845–7850. sediments. Nature 393, 567–569.
Faik, F., Belfoul, M.A., Bouabdelli, M., Hassenforder, B., 2001. The structures of the Late Jensen, S., Saylor, B.Z., Gehling, J.G., Germs, G.J., 2000. Complex trace fossils from the
Neoproterozoic and Early Palæozoic cover of the Tata area, western Anti-Atlas, terminal Proterozoic of Namibia. Geology 28, 143–146.
Morocco: polyphased deformation or basement/cover interactions during the Jensen, S., 2003. The Proterozoic and earliest trace fossil record; patterns, problems and
Variscan orogeny? J. Afr. Earth Sc. 32, 765–776. perspectives. Integr. Comp. Biol. 43, 219–228.
Fitch, A.A., 1850. A historical, topographical, and agricultural survey of the County of Karaoui, B., Breitkreuz, C., Mahmoudi, A., Youbi, N., Hofmann, M., Gärtner, A.,
Washington. Trans. New York Agric. Soc. 9, 753–944. Linnemann, U., 2015. U-Pb zircon ages from volcanic and sedimentary rocks of the
Føyn, S., Glaessner, M.F., 1979. Platysolenites, other animal fossils, and the Precambrian- Ediacaran Bas Draâ inlier (Anti-Atlas Morocco): chronostratigraphic and provenance
Cambrian transition in Norway. Nor. Geol. Tidsskr. 59, 25–46. implications. Precambr. Res. 263, 43–58.
Frey, R. W., Howard, J. D., 1985. Upper Cretaceous trace fossils, Book cliffs of Utah: A Kaufman, A.J., Knoll, A.H., 1995. Neoproterozoic variations in the C-isotopic
field guide. In: Depositinal facies of the Castlegate and Blackhawk Formations, Book composition of seawater: stratigraphic and biogeochemical implications. Precambr.
Cliffs, Eastern Utah. Soc. Econ. Paleontol. Mineralog. Rocky Mountain Section. Res. 73, 27–49.
115–152. Kesidis, G., Slater, B.J., Jensen, S., Budd, G.E., 2019. Caught in the act: priapulid
Gasquet, D., Ennih, N., Liégéois, J.P., Soulaimani, A., Michard, A., 2008. The PanAfrican burrowers in early Cambrian substrates. Proc. R. Soc. B 286 (1894), 20182505.
belt. In: Michard, A., Chalouan, A., Saddiqi, O., de Lamotte, D. (Eds.), Continental Kimura, H., Matsumoto, R., Kakuwa, Y., Hamdi, B., Zibaseresht, H., 1997. The
Evolution: The Geology of Morocco. Structure, Stratigraphy, and Tectonics of the VendianCambrian δ13C record, North Iran: evidence for overturning of the ocean
African-Atlantic-Mediterranean Triple Junction. Lect. Not. Earth Sci. 116, 33–64. before the Cambrian Explosion. Earth Planet. Sci. Lett. 147, E1–E7.
Gehling, J.G., Jensen, S., Droser, M.L., Myrow, P.M., Narbonne, G.M., 2001. Burrowing Kimura, H., Watanabe, Y., 2001. Ocean anoxia at the Precambrian-Cambrian boundary.
below the basal Cambrian GSSP, fortune head, Newfoundland. Geol. Mag. 138, Geology 29, 995–998.
213–218. Knaust, D., 2017. Atlas of Trace Fossils in Well Core. Appearance, Taxonomy
Germs, G.J., 1995. The Neoproterozoic of southwestern Africa, with emphasis on andInterpretation. Springer. 219 pp.
platform stratigraphy and paleontology. Precambr. Res. 73, 137–151. Knoll, A.H., Carroll, S.B., 1999. Early animal evolution: emerging views from
Geyer, G., 2005. The Fish River Subgroup in Namibia: stratigraphy, depositional comparative biology and geology. Science 284, 2129–2137.
environments, and the Proterozoic-Cambrian boundary problem revisited. Geol. Kouchinsky, A., Bengtson, S., Landing, E., Steiner, M., Vendrasco, M., Ziegler, K., 2017.
Mag. Terreneuvian stratigraphy and faunas from the Anabar Uplift, Siberia. Acta
Geyer, G., Uchman, A., 1995. Ichnofossil assemblages from the Nama Group Palaeontol. Pol. 62, 311–440.
(Neoproterozoic-Lower Cambrian) in Namibia and the Proterozoic-Cambrian Książkiewicz, M., 1977. Trace fossils in the flysch of the Polish Carpathians. Palaeontol.
boundary problem revisited. Beringeria Special Issue 2, 175–202. Pol. 36, 1–208.
Gibb, S., Pemberton, S.G., Chatterton, B.D.E., 2017. Arthropod Trace Fossils of the Upper Landing, E., 1994. Precambrian-Cambrian boundary global stratotype ratified and a new
Lower Cambrian Gog Group, Southern Rocky Mountains of Canada. Ichnos 24, perspective of Cambrian time. Geology 22, 179–182.
91–123. Landing, E., Geyer, G., Brasier, M.D., Bowring, S.A., 2013. Cambrian evolutionary
Hagadorn, J.W., Fedo, C.M., Waggoner, B.M., 2000. Early Cambrian Ediacaran type radiation: context, correlation, and chronostratigraphy–overcoming deficiencies of
fossils from California. J. Paleo. 74, 731–740. the first appearance datum (FAD) concept. Earth Sci. Rev. 123, 133–172.
Hall, J., 1847–1852. Palaeontology of New York: v. 1, 338 pp. (1847); v. 2: 362 pp. Letsch, D., Large, S.J.E., Bernasconi, S.M., Klug, C., Blattmann, T.M., Winkler, W., von
(1852). New York (Albany). Quadt, A., 2019. Northwest Africa’s Ediacaran to early Cambrian fossil record, its

13
A. Azizi et al. Precambrian Research 399 (2023) 107227

oldest metazoans and age constraints for the basal Taroudant Group (Morocco). Pemberton, S.G., Frey, R.W., Bromley, R.G., 1988. The ichnotaxonomy of Conostichus
Precambr. Res. 320, 438–453. and other plug-shaped ichnofossils. Can. Jour. Earth Sci. 25, 866–892.
Li, D.A., Ling, H.F., Jiang, S.Y., Pan, J.Y., Chen, Y.Q., Cai, Y.F., Feng, H.Z., 2009. New Pervesler, P., Uchman, A., Hohenegger, J., Dominici, S., 2011. Ichnological record of
carbon isotope stratigraphy of the Ediacaran-Cambrian boundary interval from SW environmental changes in Early Quaternary (Gelasian–Calabrian) marine deposits of
China: implications for global correlation. Geol. Mag. 146, 465–484. the Stirone Section, Northern Italy. PALAIOS 26, 578–593.
Linnemann, U., Ovtcharova, M., Schaltegger, U., Gärtner, A., Hautmann, M., Geyer, G., Phillips, C.h., Mcilroy, D., Elliott, T., 2011. Ichnological characterization of Eocene/
Smith, J., 2019. New high-resolution age data from the Ediacaran-Cambrian Oligocene turbidites from the Grès d’Annet Basin, French Alps, SE France.
boundary indicate rapid, ecologically driven onset of the Cambrian explosion. Terra Palaeogeogr. Palaeoclimatol. Palaeoecol. 300, 67–83.
Nova 31, 49–58. Pickerill, R.K., Peel, J.S., 1990. Trace fossils from the Lower Cambrian Bastion Formation
Loyd, S.J., Marenco, P.J., Hagadorn, J.W., Lyons, T.W., Kaufman, A.J., Sour-Tovar, F., of North-East Greenland. Rapp. Grønl. Geol. Unders. 147, 5–43.
Corsetti, F.A., 2012. Sustained low marine sulfate concentrations from the Rodríguez-Tovar, F.J., Uchman, A., 2004. Ichnotaxonomic analysis of the Cretaceous/
Neoproterozoic to the Cambrian: insights from carbonates of northwestern Mexico Palaeogene boundary interval in the Agost section, south-east Spain. Cretac. Res. 25,
and eastern California. Earth Planet. Sci. Lett. 339, 79–94. 635–647.
Loyd, S.J., Marenco, P.J., Hagadorn, J.W., Lyons, T.W., Kaufman, A.J., Sour-Tovar, F., Saadi, M., Hilali, E.A., Bensaïd, M., Boudda, A., Dahmani, M., 1985. Carte géologique du
Corsetti, F.A., 2013. Local 34S variability in ~580 Ma carbonates of northwestern Maroc 1/1000000. (Edition du Service géologique du Maroc).
Mexico and the Neoproterozoic marine sulfate reservoir. Precambr. Res. 224, Sarkar, S., Ghosh, S.K., Chakraborty, C., 2009. Ichnology of a Late Palaeozoic ice-
551–569. marginal shallow marine succession: Talchir Formation, Satpura Gondwana basin,
Maloof, A.C., Schrag, D.P., Crowley, J.L., Bowring, S.A., 2005. An expanded record of central India. Palaeogeogr. Palaeoclimatol. Palaeoecol. 283, 28–45.
Early Cambrian carbon cycling for the Anti-Atlas margin, Morocco. Can. J. Earth. Schmitt, M., 1979. The section of Tiout (Precambrian/Cambrian boundary beds, Anti-
Sci. 42, 2195–2216. Atlas, Morocco): Stromatolites and their biostratigraphy. Arb. Paläont. Inst.
Maloof, A.C., Porter, S.M., Moore, J.L., Dudas, F.Ö., Bowring, S.A., Higgins, J.A., Fike, D. Würzburg. 2, 1–188.
A., Eddy, M.P., 2010. The earliest Cambrian record of animals and ocean Schmitt, M., Monninger, W., 1977. Stromatolites and Thrombolites in Precambrian/
geochemical change. Geol. Soc. Am. Bull. 122, 1731–1774. Cambrian Boundary Beds of the Anti-Atlas, Morocco: Preliminary Results. In:
Mángano, M.G., Buatois, L.A., 2014. Decoupling of body-plan diversification and Flügel, E. (Ed.), Fossil Algae. Springer, Berlin, pp. 80–85.
ecological structuring during the Ediacaran-Cambrian transition: evolutionary and Scott, J.J., Renaut, R.W., Buatois, L.A., Owen, R.B., 2009. Biogenic structures in exhumed
geobiological feedbacks. Proc. Royal Soc. B 281, 20140038. surfaces around saline lakes: an example from Lake Bogoria, Kenya Rift Valley.
Mángano, M.G., Buatois, L.A., 2016. The Cambrian Explosion. In: Mángano, M.G., Palaeogeogr. Palaeoclimatol. Palaeoecol. 272, 176–198.
Buatois, L.A. (Eds.), The Trace-Fossil Record of Major Evolutionary Events. Volume Sdzuy, K., 1978. The Precambrian-Cambrian boundary beds in Morocco (Preliminary
1: Precambrian and Paleozoic, Volume 1. Springer, Netherlands, Dordrecht, Report). Geol. Mag. 115, 83–94.
pp. 73–126. Seilacher, A., 1999. Biomat-related lifestyles in the Precambrian. PALAIOS 14, 86–93.
Mángano, M.G., Buatois, L.A., 2017. The Cambrian revolutions: trace-fossil record, Seilacher, A., 2007. Trace Fossil Analysis. Springer 1–226.
timing, links and geobiological impact. Earth Sci. Rev. 173, 96–108. Seilacher, A., Buatois, L.A., Mangano, M.G., 2005. Trace fossils in the Ediacaran
Männil, R. M., 1966. O vertikalnykh norkakh zaryvaniya v Ordovikskikh izvestnyakakh Cambrian transition; behavioral diversification, ecological turnover and
Pribaltiki. [A small vertically excavated cavity in Baltic Ordovician limestone.] environmental shift. Palaeogeogr, Palaeoclimat, Palaeoecol. 227, 323–356.
Akademiya Nauk SSSR. Paleontologicheskiy Institut, St. Petersburg, pp. 200–207. Shahkarami, S., Mangano, M.G., Buatois, L.A., 2017. Ichnostratigraphy of the
Marusin, V. V., Kolesnikova, A. A., Kochnev, B. B., Kuznetsov, N. B., Pokrovsky, B. G., Ediacaran–Cambrian boundary: new insights on lower Cambrian biozonations from
Romanyuk, T. V., Karlova, G.A., Rud’ko, S.V., Shatsillo, A.V., Dubenskiy. A.S., the Soltanieh Formation of northern Iran. J. Paleontol. 91, 1178–1198.
Sheshukov, V.S., Lyapunov, S. M., 2021. Detrital zircon age and biostratigraphic and Sharma, M., Pandey, S.K., Ahmad, S., Kumar, K., Ansari, A.H., 2018. Observations on the
chemostratigraphic constraints on the Ediacaran–Cambrian transitional interval in ichnospecies Monomorphichnus multilineatus from the Nagaur Sandstone
the Irkutsk Cis–Sayans Uplift, southwestern Siberian Platform. Geol. Mag, 158, (Cambrian Series 2-Stage 4), Marwar Supergroup, India. J. Earth Syst. Sci 127, 1–31.
1156–1172. Smith, E.F., Macdonald, F.A., Petach, T.A., Bold, U., Schrag, D.P., 2016. Integrated
McIlroy, D., Brasier, M.D., 2017. Ichnological evidence for the Cambrian explosion in the stratigraphic, geochemical, and paleontological late Ediacaran to early Cambrian
Ediacaran to Cambrian succession of Tanafjord, Finnmark, northern Norway. Geol. records from southwestern Mongolia. GSA Bull. 128, 442–468.
Soc. Lond. Spec. Publ. 448, 351–368. Smith, E.F., Nelson, L.L., Tweedt, S.M., Zeng, H., Workman, J.B., 2017. A cosmopolitan
Meyer, M., Xiao, S., Gill, B.C., Schiffbauer, J.D., Chen, Z., Zhou, C., Yuan, X., 2014. late Ediacaran biotic assemblage: new fossils from Nevada and Namibia support a
Interactions between Ediacaran animals and microbial mats: insights from Lamonte global biostratigraphic link. Proc. Royal Soc. B 284, 20170934.
trevallis, a new trace fossil from the Denying Formation of South China. Palaeogeogr. Soulaimani, A., Ouanaimi, H., Saddiqi, O., Baidder, L., Michard, A., 2018. The Anti-
Palaeoclimatol. Palaeoecol. 396, 62–74. AtlasPan-African Belt (Morocco): Overview and Pending Questions. C. R. Geosci.
Nance, R. D., Evans, D. A., Murphy, J. B., 2022. Pannotia: To be or not to be?. Earth. Sci. 350, 279–288.
Rev, 104128. Sour-Tovar, F., Hagadorn, J.W., Huitrón-Rubio, T., 2007. Ediacaran and Cambrian index
Narbonne, G.M., 2004. Modular construction of early Ediacaran complex life forms. fossils from Sonora, Mexico. Palaeontology 50, 169–175.
Science 305, 1141–1144. Stammeier, J.A., Hippler, D., Nebel, O., Leis, A., Grengg, C., Mittermayr, F., Kasemann, S.
Narbonne, G.M., 2005. The Ediacara Biota: neoproterozoic origin of animals and their A., Dietzel, M., 2019. Radiogenic Sr and Stable C and O isotopes across Precambrian-
ecosystems. Annu. Rev. Earth Planet. Sci. 33, 421–442. Cambrian transition in marine carbonatic phosphorites of Malyi Karatau
Narbonne, G.M., Aitken, J.D., 1990. Ediacaran fossils from the Sekwi Brook area, (Kazakhstan) – Implications for paleo-environmental change. Geochem. Geophys.
Mackenzie Mountains, northwestern Canada. Palaeontology 33, 945–980. Geosyst. 20, 3–23.
Narbonne, G.M., Myrow, P., Landing, E., Anderson, M.M., 1987. A candidate stratotype Steiner, M., Yang, B., Hohl, S., Zhang, L., Chang, S., 2020. Cambrian small skeletal fossil
for the Precambrian-Cambrian boundary, Fortune Head, Burin Peninsula, and carbon isotope records of the southern Huangling Anticline, Hubei (China) and
southeastern Newfoundland, Canada. J. Earth Sci. 24, 1277–1293. implications for chemostratigraphy of the Yangtze Platform. Palaeogeogr.
Narbonne, G.M., Kaufman, A.J., Knoll, A.H., 1994. Integrated chemostratigraphy and Palaeoclimatol. Palaeoecol. 554, 109817 https://doi.org/10.1016/j.
biostratigraphy of the Windermere Supergroup, northwestern Canada: implications palaeo.2020.109817.
for Neoproterozoic correlations and the early evolution of animals. Geol. Soc. Am. Tarhan, L.G., Droser, M.L., Planavsky, N.J., Johnston, D.T., 2015. Protracted
Bull. 106, 1281–1292. development of bioturbation through the early Palaeozoic Era. Nat. Geosci.
Nelson, L.L., Crowley, J.L., Smith, E.F., Schwartz, D.M., Hodgin, E.B., Schmitz, M.D., 865–886.
2023. Cambrian explosion condensed: High-precision geochronology of the lower Thomas, R.J., Chevallier, L.P., Gresse, P.G., Harmer, R.E., Eglington, B.M., Armstrong, R.
Wood Canyon Formation, Nevada. Proc. Natl. Acad. Sci. 120, 2301478120. A., de Beer, C.H., de Beer, C.H., Martini, J.E.J., de Kock, G.S., Macey, P.H.,
Nelson, L. L., Ramezani, J., Almond, J. E., Darroch, S. A., Taylor, W. L., Brenner, D. C., Ingram, B.A., 2002. Precambrian evolution of the Sirwa window, Anti-Atlas orogeny,
Brenner, R., Madison., T., Smith, E. F., 2022. Pushing the boundary: A calibrated Morocco. Precambr. Res. 118, 1–57.
Ediacaran-Cambrian stratigraphic record from the Nama Group in northwestern Thomas, R.J., Fekkak, A., Ennih, N., Errami, E., Loughlin, S.C., Gresse, P.G., Chevallier, L.
Republic of South Africa. Earth Planet. Sci. Lett. 580, 117396. P., Liégeois, J.P., 2004. A new lithostratigraphic framework for the Anti-Atlas
Nicholson, H.A., 1873. Contributions to the study of the errant annelids of the older Orogen, Morocco. Precambr. Res. 39, 217–226.
Paleozoic rocks. Proc. Royal Soc. Lond. 21, 288–290. Triantafyllou, A., Berger, J., Baele, J.M., Diot, H., Ennih, N., Plissart, G., Monnier, C.,
O’Connor, E.A., 2010. Geology of the Drâa, Kerdous, and Boumalne Districts, Anti-Atlas, Watlet, A., Bruguier, O., Spagna, P., Vandycke, S., 2016. The
Morocco. Technical Report IR/10/072. British Geological Survey, Keyworth, Tachakoucht–Iriri–Tourtit arc complex (Moroccan Anti-Atlas): Neoproterozoic
Nottingham, England, pp. 310. records of polyphased subduction-accretion dynamics during the Pan-African
O’Neil, G.R., Tackett, L.S., Meyer, M., 2020. Petrographic evidence for Ediacaran orogeny. J. Geodyn. 96, 81–103.
microbial mat-targeted behaviors from the great basin, United States. Precambr. Res. Turk, K.A., Maloney, K.M., Laflamme, M., Darroch, S.A., 2022. Paleontology and
345, 105768. ichnology of the late Ediacaran Nasep-Huns transition (Nama Group, southern
Osgood, R.G., 1970. Trace fossils of the Cincinnati area. Paleontogr. Am. 6, 281–438. Namibia). J. Paleo. 96, 753–769.
Peel, J.S., 2010. Articulated hyoliths and other fossils from the Sirius Passet Lagerstaette Uchman, A., Kazakauskas, V., Gaigalas, A., 2009. Trace fossils from Late Pleistocene
(early Cambrian) of North Greenland. Bull. Geosci. 8, 385–394. lacustrine varved sediments in eastern Lithuania. Palaeogeogr. Palaeoclimatol.
Pelechaty, S.M., Grotzinger, J.P., Kashirtsev, V.A., Zhernovsky, V.P., 1996. Palaeoecol. 272, 199–211.
Chemostratigraphic and sequence stratigraphic constraints on Vendian-Cambrian Vannier, J., Calandra, I., Gaillard, C., Żylińska, A., 2010. Priapulid worms: Pioneer
basin dynamics, Northeast Siberian Craton. J. Geol. 104, 543–563. horizontal burrowers at the Precambrian-Cambrian boundary. Geology 38, 711–714.
Pemberton, S.G., Frey, R.W., 1982. Trace fossil nomenclature and the Planolites- Vinn, O., Wilson, M.A., Toom, U., 2015. Distribution of Conichnus and Amphorichnus in
Palaeophycus dilemma. J. Paleo. 56, 843–881. the Lower Paleozoic of Estonia (Baltica). Carnets de Geologie. 15, 269–278.

14
A. Azizi et al. Precambrian Research 399 (2023) 107227

Walsh, G.J., Aleinikoff, J.N., Benziane, F., Yazidi, A., Armstrong, T.R., 2002. U-Pb zircon Youbi, N., Kouyate, D., Soderlund, U., Ernst, R., Soulaimani, A., Hafid, A., Ikenne, M., El
geochronology of the Paleoproterozic Tagragra de Tata inlier and its Neoproterozoic Bahat, A., Bertrand, H., Rkha Chaham, K., Ben Abbou, M., Mortaji, A., El Ghorfi, M.,
cover, western Anti-Atlas, Morocco. Precambr. Res. 117, 1–20. Zouhair, M., El Janati, M., 2013. The 1750 ma magmatic event of the west african
Walsh, G.J., Benziane, J.N., Aleinikoff, F., Harrison, R.W., Yazidi, A., Burton, W.C., craton (Anti-Atlas, Morocco). Precambr. Res. 236, 106–123.
Quick, J.E., Saadane, A., 2012. Neoproterozoic tectonic evolution of the Jebel Zhang, J., Li, G., Zhou, C., Zhu, M., Yu, Z., 1997. Carbon isotope profiles and their
Saghro and Bou Azzer-El Graara inliers, eastern and central Anti-Atlas, Morocco. correlation across the Neoproterozoic-Cambrian boundary interval on the Yangtze
Precambr. Res. 216, 23–62. Platform, China. Bull. Natl. Museum Nat. Sci. 10, 107–116.Platform, China. Bull.
Weber, B., Steiner, M., Evseev, S., Yergaliev, G., 2013. First report of a Meishucun-type Natl. Museum Nat. Sci. 10, 107–116.
early Cambrian (Stage 2) ichnofauna from the Malyi Karatau area (SE Kazakhstan): Zhu, M., Yang, A., Yuan, J., Li, G., Zhang, J., Zhao, F., Ahn, S.Y., Miao, L., 2019.
palaeoichnological, palaeoecological and palaeogeographical implications. Cambrian integrative stratigraphy and timescale of China. Sci. China Earth Sci. 62,
Palaeogeogr. Palaeoclimatol. Palaeoecol. 392, 209–231. 25–60.

15

You might also like