<<

ÔØ ÅÒÙ×Ö ÔØ

Where did ancient carbonate mounds grow – in bathyal depths or in shallow shelf waters?

Dierk Hebbeln, Elias Samankassou

PII: S0012-8252(15)00043-4 DOI: doi: 10.1016/j.earscirev.2015.03.001 Reference: EARTH 2094

To appear in: Earth Science Reviews

Received date: 15 September 2014 Revised date: 3 March 2015 Accepted date: 6 March 2015

Please cite this article as: Hebbeln, Dierk, Samankassou, Elias, Where did ancient car- bonate mounds grow – in bathyal depths or in shallow shelf waters?, Earth Science Reviews (2015), doi: 10.1016/j.earscirev.2015.03.001

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain. ACCEPTED MANUSCRIPT

Where did ancient carbonate mounds grow – in bathyal depths or in shallow shelf waters?

Dierk Hebbelna & Elias Samankassoub a MARUM – Center for Marine Environmental Sciences, University of Bremen, Leobener Straße,

28359 Bremen, Germany b Section of Earth and Environmental Sciences, University of Geneva, Rue des Maraîchers 13, 1205

Geneva, Switzerland

corresponding author:

Dierk Hebbeln

Tel. +49-421-21865650

Fax +49-421-21865654 email: [email protected]

Abstract

Carbonate mounds are important and ubiquitous components of the sedimentary rock record throughout the PhanerozoicACCEPTED. Nevertheless, factors MANUSCRIPT controlling their occurrence and growth remain enigmatic, in particular because the depositional depths of such mounds are poorly constrained and it is assumed that fossil examples lack modern analogues. Factors currently used to constrain the water depth ancient mounds grew in include water energy, occurrence of specific organisms that perform photosynthesis (particularly calcareous algae) and distinct non-biogenic components. In reviewing the main criteria used to delineate water depth of formation of fossil carbonate mounds, most (if not all) criteria traditionally used to assume shallow-water settings appear to be no longer substantiated in the light of current knowledge gained from modern carbonate mounds. High energy, assumed to indicate shallow-water conditions, is ongoing in water as deep as > 1000 m. The

1

ACCEPTED MANUSCRIPT

occurrence of peloids and processes such as micritization are not diagnostic of water depth.

Furthermore, virtually none of the calcareous algae assumed to indicate the have a taxonomically granted affiliation.

Over the last years, research in the deep , in particular thanks to the deployment of new devices and techniques, provided a good database revealing a wide distribution of modern carbonate mounds. Reviewing the characteristics of ancient carbonate mounds, with emphasis on depositional settings, and confronting the criteria used with current knowledge from sub-recent and modern carbonate mounds, we conclude that for many ancient mounds a bathyal origin is (at least) as likely as a shallow-water origin. This conclusion is well supported by the sedimentary records left by these modern mounds that are astonishingly similar to those seen in Palaeozoic and carbonate mounds.

Keywords: Carbonate mounds – Waulsortian mounds – –Palaeozoic – Modern

1. Introduction

Carbonate mounds are common sedimentary features. They have been formed by a variety of organisms since the early Palaeozoic (Monty et al., 1995). Framework-building organisms such as corals, crinoids and bryozoans commonly represent important components in these mounds, ACCEPTED MANUSCRIPT although pure microbial mud mounds are also widely reported (Pratt, 1995). The structure and genesis of fossil carbonate mounds are a matter of debate, basically because true modern analogues are thought to be scarce (Wendt et al., 2001). In analogy to modern tropical coral reefs, most ancient carbonate mounds have widely been linked with ramp/shelf settings in tropical environments (Lees and Miller, 1995), although other depositional environments have been suggested (cf. review in

Pratt, 1995).

A shelf origin was supported by the assumption that high carbonate production in oligotrophic settings, common for modern coral reefs, has to rely on photoautotrophic symbionts producing

2

ACCEPTED MANUSCRIPT

energy for the framework builders (e.g., zooxanthellate scleractinian corals) (Stanton, 2006; Pomar and Hallock, 2008). However, Stanton (2006) pointed out that for ancient mounds the rate of carbonate production could not be determined confidently. Reviewing the evidence for an origin of carbonate mounds in the photic zone, Stanton (2006) concluded that for much of the Phanerozoic

(that is, at least prior to the and, perhaps, Late ) a direct photosynthetically mediated carbonate production in mounds cannot be proven and appears to be unlikely.

While the presence of photoautotrophic organisms in the fossil record of an ancient mound indeed would be a distinctive indicator for an origin within the photic zone, other criteria commonly used to infer a shallow-water origin are rather equivocal. Basically these criteria can be grouped into two lines of thought: (a) high-energy environments often invoked for ancient carbonate mounds only appear in shallow waters linked to fair weather or storm wave bases (Tsien et al., 1980; Brachert et al., 1992; Kirby and Hunt, 1996; Boulvain, 2001; Boulvain et al., 2004; Pomar and Hallock, 2008), and

(b) specific organisms or non-bioclastic components are (mostly) limited in their ecological range to shallow waters (Lees and Miller, 1985; Madi et al., 1996; cf. discussion in Monty, 1995).

In addition to tropical coral reefs, modern, actively growing carbonate mounds that usually occur in aphotic depths have a great potential also to serve as modern analogues for many ancient carbonate mounds (Mullins et al., 1981; Cairns and Stanley, 1981). Although these modern carbonate mounds have been found in inner and outer shelf settings, their vast majority has been reported from depths ACCEPTED MANUSCRIPT between 200 m and 1000 m, thus from well below the shelf break (Roberts et al., 2006). Over the last decades the application of state-of-the-art marine technologies including remotely operated vehicles

(ROVs), autonomous underwater vehicles (AUVs), and multibeam echosounders helped to detect carbonate mounds in increasingly important numbers, e.g., along the NW European margin (Fosså et al., 2005; Wheeler et al., 2007), some Mediterranean margins (Freiwald et al. 2009; Fink et al., 2013), the NW African margin (Colman et al., 2005; Foubert et al., 2008), the Florida-Hatteras margin (Paull et al., 2000; Grasmück et al., 2006), the Brazilian (Viana et al., 1998), Mexican (Hebbeln et al., 2014), and Angolan margins (Le Guilloux et al., 2009), the Canadian Pacific margin (Conway et al., 2005), and

3

ACCEPTED MANUSCRIPT

in the Great Australian Bight (James et al., 2000). All these mounds depend on skeletal material produced by heterozoans with the most important components being azooxanthellate scleractinian corals (in the following referred to as cold-water corals, CWC, e.g., Roberts et al., 2006), bryozoans

(James et al., 2000) and sponges (Conway et al., 2005). Although reaching partly back with their origins to >2.5 Ma (Kano et al., 2007), most of the CWC carbonate mounds have been hosting active coral communities in recent and/or sub-recent (i.e. <50 kyr old) times (Frank et al., 2011; Fink et al.

2013).

In contrast to such heterozoan mounds, adequate modern analogues for pure microbial mud mounds are still waiting to be discovered. However, assuming that such microbial mud mounds also form today in the deeper ocean, their presumably low-relief features might only be discovered following a further refinement of seafloor mapping technologies.

The aim of this review paper is to show that these modern carbonate mounds from well below the shelf break can serve as equally good analogues to ancient heterozoan carbonate mounds as modern shallow-water carbonate mounds do. Already in the early days of research on modern carbonate mounds, their potential role as analogues for ancient carbonate mounds has been acknowledged

(Mullins et al., 1981; Cairns and Stanley, 1981). However, whereas in the early 1980s modern carbonate mounds growing in several 100s of meters water depth might have been seen as peculiar and rare features, now their well-documented widespread occurrence (see above; Fig. 1) justifies a ACCEPTED MANUSCRIPT new approach linking the research on modern and ancient heterozoan carbonate mounds, with the latter also showing a wide geographical distribution (Fig. 1).

Whereas the dominant organisms and, consequently, the biotic interactions on such mounds have been variable over the last several hundreds of million years resulting in a large variety of individual mound types (Kiessling et al., 2002; James and Wood, 2010), all these types share a common carbonate mound appearance. This calls for a common set of processes being involved in their formation affecting both, ancient and modern carbonate mounds. Indeed, a closer look at the modern carbonate mounds occurring beyond the shelf break reveals that they share many of the

4

ACCEPTED MANUSCRIPT

characteristic features, which are commonly used to infer a shallow-water origin of ancient carbonate mounds. Consequently, the sedimentary records of modern (e.g., Dorschel et al., 2005) as well as of ancient (Brachert et al., 1992; Wendt, 1993; Pratt, 1995) carbonate mounds are astonishingly similar in size and composition (Figs. 2, 3). Demonstrating the similarities (1) of their appearance, (2) of the processes involved in their formation, and (3) of the resulting sedimentary records, this paper might stimulate an open discussion on the origin of the ancient carbonate mounds and trigger a more intense dialogue between scientists studying either ancient or modern mound systems (see also Henriet et al., 2014). By acknowledging the potential of these modern mounds, forming in the to serve as analogues for ancient mounds, such a discussion finally may result in the re-assessment of the formation depth of many ancient carbonate mounds.

Fig. 1: Distribution of cold-water corals (dots and lines in light blue, after Roberts et al., 2006) in the

Atlantic Ocean. Superimposed are examples of proven occurrences of recent (stars, mostly representing clusters of mounds - compiled from Colman et al., 2005; Fink et al., 2013; Foubert et al.,

2008, Grasmueck et al., 2006; Hebbeln et al., 2014; Le Guilloux et al., 2009, Mienis et al. 2012; 2014;

Mullins et al., 1981; Remia and Taviani, 2005; Reveillaud et al., 2008; Reyes et al., 2005; Taviani et al.,

2005; Viana et al., 1998; Wheeler et al. 2007) and of ancient carbonate mounds (triangles, compiled from compiled from Wilson, 1975; Geldsetzer et al., 1989; Flügel and Flügel-Kahler, 1992; Bourque et ACCEPTED MANUSCRIPT al., 1995; Pratt, 1995; Gutteridge et al., 1995; Lees and Miller, 1995; Madi et al., 1996; Strogen et al., 1996; Boulvain, 2001; Wendt et al., 2001, Riding, 2002; Webb, 2002). At many sites for which the occurrence of cold-water corals has been described, modern carbonate mounds still might be discovered, as for example recently along the Angolan margin (Le Guilloux et al., 2009), the Moroccan Mediterranean margin (Fink et al., 2013), and the Mexican Campeche Bank (Hebbeln et al., 2014).

5

ACCEPTED MANUSCRIPT

Fig. 2: Examples for the shape and dimensions of ancient and modern carbonate mounds. Upper panel: Province of exhumed mounds (Kess Kess mounds, Devonian) from the central area of Hamar

Laghdad Ridge, Anti-Atlas, Morocco (details on composition of these mounds are provided in

Brachert et al., 1992). Individual mounds on the image are in average 30 m high. Lower panel:

Multibeam echosounder map of cold-water coral carbonate mounds at the Moroccan Atlantic continental margin with dimensions indicated. Data have been obtained in 2014 during expedition

MSM-36 with the German RV Maria S. Merian (Hebbeln et al., unpubl. data).

Fig. 3: Comparison of the lithology of ancient and modern carbonate mounds. (a) Vertical cut-surface of sediment core GeoB 14904-2 from a Mauritanian margin cold-water coral mound (to the left) compared to a vertical outcrop surface (to the right) of a Visean carbonate mound, Moroccan Anti-

Atlas (for the setting see Wendt et al., 2001). Scale bar is 15 cm. The white rectangle indicates the zoom-in displayed in (b).

2. Some definitions

Since many years there is an ongoing discussion about a classification of ancient carbonate mounds involving, among many others, terms ranging from microbial mud mounds via skeletal carbonate build-ups to reefs (James and Bourque, 1992; Bosence and Bridges, 1995; Wood, 2001; Riding, 2002; ACCEPTED MANUSCRIPT Kiessling et al., 2002; James and Wood, 2010). The commonly used term “Waulsortian mounds” essentially refers to early Mississippian mud-rich mounds (Lees et al., 1985). As this paper is interested in the processes of mound formation, the general term “carbonate mounds” is used, in order to avoid any controversial discussions in relation to the various classification schemes.

Most of the modern carbonate mounds referred to above occur beyond the shelf break in water depths down to >1000 m (White and Dorschel, 2010), although in temperate and high latitudes they also can occur in inner and outer shelf settings (Freiwald et al., 1997; Lavaleye et al., 2009). By increasing the range of mound occurrence to beyond the shelf edge, terms such as “shallow-water

6

ACCEPTED MANUSCRIPT

mounds” and “deep-water mounds”, which are commonly used in relation to ancient carbonate mounds, might create some confusion as especially the latter term often still refers to shelf settings.

Definitions for the differentiation of such palaeo-depth estimates are manifold and often related to either the photic zone, with “deep water” still referring to within photic zone (e.g., Pratt 1995) or to hydrodynamic energy and with “deep water” still referring to above the storm- (Brachert et al., 1992; Pratt, 1995). Thus, the lower limits given for the bathymetry for ancient deep-water carbonate mounds mostly range between ~30 m and ~280 m (Lees et al., 1985; Lees and Miller,

1985; Bridges and Chapman, 1988; Monty et al., 1995; Pratt, 1995; Bourque et al., 1995; Jeffery and

Stanton, 1996). To differentiate these ancient deep-water mounds from the modern carbonate mounds occurring in much greater water depths, the latter will be assigned as “bathyal mounds” in the following, and this term also is recommended for future work.

3. Shelf versus bathyal origin of carbonate mounds

As pointed out above, indicators to allocate a shallow-water origin to ancient carbonate mounds in absence of photosynthetic organisms in their fossil record are often equivocal. In the following, the two main arguments, namely (a) indications for high-energy settings hinting to a formation above the storm/fair-weather wave base (Tsien et al, 1980; Brachert et al., 1992; Kirby and Hunt, 1996;

Boulvain, 2001; Pomar and Hallock, 2008) and (b) specific mound organisms being restricted to ACCEPTED MANUSCRIPT shallow waters (e.g., Lees and Miller, 1985, 1995; Madi et al., 1996) are critically assessed and compared to modern bathyal carbonate mound settings.

Of course, when comparing ancient with modern carbonate mounds (e.g., Fig. 3), the impact of diagenesis - primarily on the fossil record - has to be kept in mind. In his ‘diagenetic sieve’ approach,

Dullo (1990) pointed out that especially selective dissolution of biogenic components might cause a significant quantitative modification of the biotic composition preserved in the rock record. Selective dissolution (e.g., of aragonitic material) potentially might even affect the quantitative composition of a mound, but probably with no significant impact on the overall appearance of a mound. For

7

ACCEPTED MANUSCRIPT

example, over the Holocene the sedimentary record on Propeller Mound at the Irish margin contains up to 15% aragonitic coral material (Dorschel et al., 2007) and even the complete dissolution of this material would hardly affect the dimensions of this ~120 m high mound. And actually, for most of the

Paleozoic carbonate mounds referred to in the present study, dissolution is overall a minor issue

(Bathurst, 1971). However, as for the reconstruction of the palaeoenvironmental setting of mound formation predominantly the preserved, i.e. still readable (fossil) record is interpreted, diagenesis would have hardly an impact on comparisons like such done here. Nevertheless, still there is the chance/risk that diagenesis might have (irretrievably) erased clear indicators for specific settings – a common threat to all geological records.

3.1. High-energy environments

As most of the common heterozoans contributing to ancient and modern carbonate mounds do not have photoautotrophic symbionts, they have to rely on external food supply that most likely originates from primary production in the surface waters. For carbonate mounds largely formed by immobile suspension feeders as, e.g., cold-water corals, bryozoans, crinoids and sponges, the food particles raining down from the surface can also be transported laterally close to the sea floor, thereby significantly increasing the chances of these organisms to catch the food. Thus, in addition to a sufficiently high surface water production, also a vigorous bottom current regime is needed, ACCEPTED MANUSCRIPT powered either by geostrophic currents or by internal . In fact, all modern bathyal carbonate mounds characterized by a thriving mound ecosystem are intrinsically linked to a dynamic bottom water regime (see Roberts et al., 2006, for cold-water coral mounds). In situ current measurements at such sites revealed maximum current velocities of >50 cm s-1 (Dorschel et al., 2007; Mienis et al.,

2009a). Such strong currents can in various ways and at various scales interact with the mounds, from creating small sedimentary structures (e.g., ripples) to controlling the shape of a mound (e.g., elongated mounds surrounded by moats). The occurrence of such sedimentary structures in the rock

8

ACCEPTED MANUSCRIPT

record (Tosolini et al., 2012) should therefore be interpreted considering also the existence of strong bottom currents at bathyal depths.

3.2. Coarse layers and cross-bedding

Layers with coarse sediments indicative of reworking and/or local bioclast accumulations in fossil mounds are thought to record major storm events (e.g., event stones in the Devonian Kess-Kess mounds of Brachert et al., 1992) and, thus, a deposition above the storm wave base. Tosolini et al.

(2012) used, among other criteria, the texture indicative of high energy (grainstone) to infer a shallow depositional setting for Devonian mounds (see also Wendt et al., 2001). Furthermore, cross- stratification in a crinoidal grainstone found close to carbonate mounds in Derbyshire led Gutteridge (1995) to infer a deposition above the fair-weather wave base. In contrast, for greater depths generally rather quiet conditions are assumed (Pray, 1958).

The strong currents of often several tens of centimeters per second observed at modern bathyal carbonate mounds (Dorschel et al., 2007; Mienis et al., 2009a; Hebbeln et al., 2014) also trigger sediment reworking and bedform formation. Thus, ripples (see Fig. 4) and coarse layers indicative for reworking and/or erosion in the mound record (Dorschel et al., 2005) are common features in and around recent carbonate mounds also in several hundreds of meters water depth, well below the storm wave base. ACCEPTED MANUSCRIPT

Fig. 4: Rippled sea floor with coral fragments from the Porcupine Seabight at the Irish margin in a water depth of ~600 m. The two dots are 20 cm apart from each other. The photograph has been taken in 2010 during expedition POS400 with the German RV Poseidon with the ROV CHEROKEE

(MARUM, University of Bremen, Germany).

3.3. Fossils in growth position

9

ACCEPTED MANUSCRIPT

The scarcity of larger fossils such as corals, bryozoans and crinoids in growth position in ancient mound records often has been interpreted as a clear indicator for strong hydrodynamics linked to shallow waters (Bourque et al., 1995; Aretz and Herbig, 2003). As pointed out above, strong bottom currents also shape bathyal mounds and, especially as the involved organisms do not form massive carbonate structures (Fig. 5a), can contribute to their disintegration (possibly favoured by bioerosion that is also common in modern bathyal carbonate mounds; Beuck et al., 2010) and move (parts of) these organisms. On many CWC carbonate mounds this process results in the widespread so-called coral rubble/debris facies (Huvenne et al., 2005; Wienberg et al., 2008; see Fig. 5b). As this facies is finally also preserved within the sediments, the sedimentary record found on the bathyal mounds usually reveals abundant coral fragments in a matrix of hemipelagic sediments (Figs. 5c, 3) with only rare large fossil fragments that might be interpreted as still being in growth position. Actually, the appearance of these modern sedimentary records is astonishingly similar to those of many ancient mounds (Fig. 3).

Fig. 5: Development from the living coral thicket towards the sedimentary record. (a) Living Lopehlia pertusa framework (in pink) growing on dead, but still intact framework (grey to white) in ~640 m water depth at Rockall Bank, NE Atlantic. The entire thicket reaches ca. 1 m in height. The photograph has been taken in 2004 during expedition M61-3 with the German RV Meteor with the ACCEPTED MANUSCRIPT ROV QUEST (MARUM, University of Bremen, Germany). (b) Coral rubble facies in the Porcupine Seabight at the Irish margin in ~750 m water depth (width of the image: ~ 30 cm). The photograph has been taken in 2010 during expedition POS400 with the German RV Poseidon with the ROV

CHEROKEE (MARUM, University of Bremen, Germany). (c) Computer-tomograph (CT) image of sediment core GeoB 11569-2 collected from a Mauritanian margin cold-water coral mound (width:

12 cm) showing that generally disintegrated coral fragments (visualized here by CT technique without surrounding sediment) form the mound facies.

10

ACCEPTED MANUSCRIPT

Thus, considering the composition of modern mounds consisting mainly of fragmented fossils (e.g., coral debris), the discussion as to which degree organisms occurring within mounds (particularly

Waulsortian mounds) are in growth position (cf. Ahr and Stanton, 1994) is not straightforward.

Crinoids and bryozoans, the main skeletal components in early to mid-Palaeozoic mounds (Pray,

1958; Lees and Conil, 1980; Lees et al., 1985; cf. Wood, 1998), can be even more delicate than the scleractinian corals living on modern bathyal mounds. Thus, crinoids and bryozoans often are even more prone to disintegration resulting in various degrees of preservation and probably in even less specimens having the potential to be preserved in growth position.

3.4. Episodic mound growth

Vertical changes in facies and grain assemblages are phenomena widely reported from, and often used to characterize, Waulsortian mounds (Lees, 1964; Lees et al., 1985; Lees and Miller, 1985;

Murphy, 1988; Bourque et al., 1995; Jeffery and Stanton, 1996; Madi et al., 1996). Kirby and Hunt

(1996) interpreted hiatal surfaces found within the Lower Carboniferous Muleshoe Mound in New

Mexico, USA, to indicate episodic growth. This development was interpreted to reflect the alternation of relatively well-oxygenated and high-energy conditions representative of shallow water conditions and periods of mound crises, triggered by anoxic bottom waters (Kirby and Hunt, 1996;

Stanton et al., 2000). Such an episodic growth has also been inferred for Carboniferous carbonate ACCEPTED MANUSCRIPT mounds in northern Africa based on the presence of recurring, internal packages (growth phases?) interpreted as having been induced by sea-level fluctuations (Bourque et al., 1995; Madi et al., 1996).

Vertical changes in the composition of fossil carbonate mounds are often interpreted to be indicative of sea-level fluctuations e.g., by affecting the position of the fair-weather/storm wave base (Wendt,

1993; Lees et al., 1995; Bourque et al., 1995; Boulvain et al., 2004).

Episodic growth is also common in modern bathyal carbonate mounds, e.g., triggered by changing hydrodynamics induced by climate change (Dorschel et al., 2005), changed productivity conditions

(Wienberg et al., 2010), or by temporary sub-oxic bottom waters (Fink et al., 2012). As a result of this

11

ACCEPTED MANUSCRIPT

episodic growth, modern mounds are often characterized by internal hiatal surfaces (Dorschel et al.,

2005; Eisele et al., 2008; Mienis et al., 2009b; van der Land et al., 2014), which resemble, for example, those described for the Muleshoe Mound (Kirby and Hunt, 1996).

3.5. Mound shape

According to the differentiation between mounds having a flat top and cone-shaped mounds introduced by Schlager (2003) and as recently applied by Berra et al. (2012), flattened mound tops are caused by erosion inevitably indicating that the mounds reached at least the storm wave base, whereas cone-shaped mounds are common in quiet and deep waters. In addition, morphologic features such as elongated mound shapes and current scours around them have been interpreted to reflect high-energy, rather shallow settings near the fair-weather wave base (Kirby and Hunt, 1996).

However, investigations of modern CWC carbonate mounds off reveal that these mounds underwent significant cyclic erosion at the turn from glacial to interglacial conditions when bottom currents in approx. 800 m water depth picked up from sluggish glacial to erosive interglacial levels

(Dorschel et al., 2005). Due to such erosive settings, also bathyal carbonate mounds often have flattened tops (Fig. 6).

Fig. 6: Detailed bathymetric 3D-map for Propeller Mound, an actively growing cold-water coral ACCEPTED MANUSCRIPT carbonate mound in ~800 m water depth at the Irish continental margin (Huvenne et al., 2005). The map clearly shows the flattened top of the mound (map courtesy by B. Dorschel).

Furthermore, as the bathyal carbonate mounds interact with a dynamic bottom water regime, moats commonly form around these mounds features (De Mol et al., 2002; Grasmueck et al., 2007; Fink et al., 2013, see also Fig. 6). In places where modern bathyal mounds are located in unidirectional flow regimes, elongated mound shapes are typical features most likely resulting from the preferred growth of the main suspension feeders towards the currents (Mullins et al. 1981).

12

ACCEPTED MANUSCRIPT

3.6. Organisms and non-biogenic components as bathymetry indicators

Crinoids and fenestellid bryozoans are the main bioclastic components in Waulsortian mounds in

Belgium as originally defined by Lees and Miller (1985). Sponges (mostly hexactinellids) were reported as important components in comparable mounds elsewhere (e.g., Madi et al., 1996). The proportion of the latter, having a lower preservation potential, was thought to be underestimated in earlier descriptions (Bourque and Gignac, 1983, 1986; Bourque and Boulvain, 1993). Other, volumetrically less important bioclastic components include corals, brachiopods, trilobites, ostracodes, various foraminifers and “problematic algae” (Bridges and Chapman, 1988; Brachert,

1991; Brachert et al., 1992; Jeffery and Stanton, 1996). Some of these organisms (in particular green algae) are used for bathymetric estimates for mound depositional environments, along with non- bioclastic components such as peloids and post-depositional features such as micritization (Lees and

Miller, 1985, 1995; Bridges and Chapman, 1988).

Lees et al. (1985) and Lees and Miller (1985, 1995) distinguished four component assemblages (A at the bottom to D at the top) in Waulsortian mounds. These assemblages were interpreted to represent stratigraphic phases tied to water depth. A key factor hereby is the delineation of the photic zone. For instance, Lees and Miller (1985, 1995) linked the lower limit of the photic zone to the deepest occurrence of micritization, which is assumed to be caused by algae and to occur down ACCEPTED MANUSCRIPT to a maximum water depth of 220 m. This depth limit is subsequently used to estimate the bathymetric range of other organisms. As in this case study multilocular foraminifers occur stratigraphically 30 m below the assumed lower limit of micritization, a maximum depth for the occurrence of multilocular foraminifers at 250 m water depth has been inferred. In the same line of reasoning, sponge spicules, found stratigraphically 60 m below the lowest level of micritization, are used to delineate the 280 m isobath.

This method appears questionable in many aspects. Micritization is not exclusively caused by algae as pointed out in Jeffery and Stanton (1996) and can hardly be a reliable criterion for bathymetry.

13

ACCEPTED MANUSCRIPT

Indeed, micritization is reported from settings as deep as 871m (Hook et al., 1984). Relative water depth deduced from components relative to the occurrence of micritization is thus arbitrary and, moreover, does not take into account possible displacement of components and post-depositional processes (see discussion in Jeffery and Stanton, 1996; Flügel, 2004).

Other critical components having been used for bathymetric reconstructions are peloids in geopetal fillings of voids, with the lower limit of occurrence estimated at depths around 250 m (Lees and

Miller, 1985, 1995). However, peloids can occur in virtually all settings and, thus, are not indicative for water depth (see review in Flügel, 2004). For instance, peloids including those occurring in geopetal infillings are widespread in slope rocks in the northern Bahamas, at depths ranging from

200 to 800 m (Wilber and Neumann, 1993).

Furthermore, and possibly most important, none of the presumed red and green algae occurring in

Palaeozoic mounds (problematic forms such as Fasciella, Aphralysia, Rothpletzella, Wetherella;

Wendt et al., 2001) used as indicative of a setting in the photic zone are unequivocally identified as algae, i.e. the position and taxonomic affinity of these fossils are uncertain (see Madi et al., 1996).

Fasciella is considered a red alga by Madi et al. (1996) and as Calcifolliida and, thus, of uncertain affinity (Phylum, Class and Order incertae sedis) by Vachard et al. (2004). The systematic position of

Aphralysia is controversial, ranging from foraminifer (Belka, 1981) to cyanobacterium (Vachard and

Aretz, 2004). Rothpletzella is considered as a cyanobacterium (cf. Riding, 1991) or as a calcimicrobe ACCEPTED MANUSCRIPT (Webb, 1996, 2001; Nose et al., 2006); thus, its occurrence is not necessarily diagnostic for the photic zone. The occurrence of Rothpletzella in deep-water slopes (Webb, 2001) compromises their attribution to green algae (Wray, 1972), and even to cyanobactria. The taxonomic position of

Wetherella is also uncertain, regarded as foraminifer (Wood, 1948), alga, cyanobacterium or microproblematicum (cf. Riding and Soja, 1993).

Forms ascribed to green algae by Jeffery and Stanton (1996), based on similarities in morphology, are poorly preserved and lack unequivocal taxonomic determination. All specimens illustrated are strongly fragmented. Sphaerinvia, a taxon identified as charophyte (Mamet, 1991; Mamet and Preat,

14

ACCEPTED MANUSCRIPT

2013), was found in situ in a deep, aphotic marine environment, precluding an algal affinity (Jeffery and Stanton, 1996).

4. Further aspects of carbonate mound formation

4.1. Mound and off-mound stratigraphic relationships

Often a generic link between off-mound sediments and mound facies is drawn, to increase the information about the palaeoenvironmental setting favoring mound growth (Lees et al., 1985;

Stanton et al. 2000) and to argue for the autochthonous production of carbonate on mounds embedded in siliciclastic off-mound sediments (Bosence and Bridges, 1995; Monty et al., 1995).

Although the resolution provided by biostratigraphical studies of ancient mounds is usually too low to allow any distinction of age differences between both facies, the “mound-concept”, presuming the mound rising above the surrounding seafloor (Pray, 1958; Lees et al., 1985; Meyer et al., 1995), requires a chronostratigraphic offset between on- and off-mound facies being preserved side-by- side. For the modern bathyal mounds, such offsets can be determined. The best-studied example is

Challenger Mound at the Irish margin that has been drilled during IODP Leg 307 (Ferdelman et al.,

2006). This mound grew for more than one million years to a height of ~130 m, while non-deposition or even erosion prevailed in surrounding areas, highlighting the potential of a carbonate mound to create its own depositional setting, to baffle background sediment effectively, and to create large ACCEPTED MANUSCRIPT stratigraphic offsets between mound facies and adjoining off-mound facies (Titschack et al., 2009). Furthermore, on- and off-mound facies can document very different environmental settings caused by cyclic switches of the main deposition between both settings. For instance, at and around

Propeller Mound at the Irish margin a predominant preservation of glacial sediments in the off- mound facies and of interstadial to interglacial sediments in the mound facies has been described

(Dorschel et al., 2005). Due to the autochthonous carbonate production but also to the enhanced allochthonous, pelagic carbonate sedimentation during interglacials, the mound facies contains on average 50% carbonate compared to <25% in the glacial off-mound facies (Dorschel et al., 2007).

15

ACCEPTED MANUSCRIPT

Thus, by revealing the linkages between on- and off-mound facies, especially in a stratigraphic sense

(ranging from a principal chronostratigraphic offset to alternating deposition reflecting different environmental forcings), studies on modern, bathyal carbonate mounds might trigger new approaches in interpreting ancient carbonate mounds in relation to their surrounding sediments.

The higher volume of mud in cores of fossil mounds compared to intermound/off-mound facies

(Samankassou et al., 2013) remained enigmatic, because the former was supposed to rise topographically higher (assumedly into the zone of high energy where mud should be removed) than the latter (assumedly lower energy where mud can be accumulated) (Pray, 1958). Also this apparent contradiction might become understandable when, for the interpretation of on- vs. off-mound sediments, the possible impacts of (a) (large) chronostratigraphic offsets, (b) different environmental settings, and (c) microenvironments on the mounds where fine sediments might be baffled within a biogenic structural framework are considered. Such factors would affect shallow and bathyal mounds in a very similar way.

Furthermore, especially in a shallow-water setting, bathymetric changes induced by mound growth are expected to be recorded in vertically superimposed facies that are also observed next to each other spatially, following Walther’s Law (Walther, 1894). However, facies architecture in examples of fossil sequences of mounds, e.g., in and around the Devonian mounds of Morocco, fails to show prograding sequences (Fig. 7; Wendt, 1993) that would be indicative of a progressive shift of ACCEPTED MANUSCRIPT environment through time. This observation is in agreement with results of Brachert et al. (1992) who pointed out the absence of obvious vertical change in the depositional types of sediment also for another Devonian mound setting in Morocco. These observations are in line with those of recent settings along the Irish continental margin, where the facies on top of the self-sustained CWC carbonate mounds (basically forming micro-environments) also cannot be traced spatially to shallower or deeper settings.

Figure 7

16

ACCEPTED MANUSCRIPT

(a) Sketch of a Middle Devonian sequence at Jebel el Oftal, Morocco, including carbonate mounds and the surrounding bedded limestone. Redrawn from Wendt (1993).

(b) Sequence of Devonian Kess Kess mounds and the surrounding bedded limestone of Hamar

Laghdad Ridge, Anti-Atlas, Morocco. Mounds seem to be amalgamated and/or aligned. As in the examples reported by Wendt (1993), the facies architecture lacks vertically superimposed facies

(mounds versus bedded limestone) indicative of prograding and/or shallowing-upwards sequences.

The mounds reach ca. 30 m in height.

4.2. Mound growth and sedimentation

In reviewing controls on the formation of Palaeozoic and Mesozoic carbonate mounds, Schmid et al.

(2001) conclude that mound development is restricted to conditions of reduced sedimentation and that high sediment input probably results in burial of mounds as mound growth might be outcompeted by background sedimentation. However, in the first step, the concept of mound formation requires mound growth to be faster than background sedimentation. Furthermore, especially for mounds hosting branching epibenthic organisms, baffling of hemipelagic sediment is a major component contributing to mound growth. Comparing average shelf and slope sedimentation rates (0.04 to 0.5 mm/yr; Seibold and Berger, 1993) with much higher growth rates of such organisms (e.g., for the branches of the cold-water coral Lophelia pertusa: 5 to 34 mm/yr; Roberts et ACCEPTED MANUSCRIPT al., 2009) indicates that burial of a growing carbonate mound by increasing sedimentation appears highly unlikely. Actually, individual modern bathyal carbonate mounds can reach growth rates of >2 mm/yr over several centuries to millennia (Frank et al., 2009; Fink et al., 2013), demonstrating the need for external sediment supply and the efficiency of sediment baffling by mound organisms.

Considering that even such extreme records mainly consist of corals not being preserved in growth position, but of mostly reworked (i.e. compacted) corals, underlines the capacity of coral growth to contribute to mound growth. Thus, burial of carbonate mounds appears to be only a consequence of

17

ACCEPTED MANUSCRIPT

the cessation of active mound growth induced by the deterioration of living conditions for the respective key organisms.

4.3. Mound belts and clusters

Ancient as well as modern carbonate mounds often occur in large numbers in clearly confined regions that are referred to as mound belts/clusters/provinces. This applies, e.g., for the Palaeozoic mounds of Morocco and Algeria (>100 in 440 km2, Brachert et al., 1992; Wendt et al., 1993, 1997,

2001; Fig. 2), Belgium (Aretz and Chevalier, 2007), the UK (Bridges and Chapman, 1988) and New

Mexico, USA (Jeffrey and Stanton, 1996) and for modern mounds off Norway (Fossa et al., 2005),

Ireland (>1000 mounds in distinct clusters between 500 m and 1500 m water depth along the Irish margin; Dorschel et al., 2010), the Bahamas (along the western slope of Great Bahamas Bank; Correa et al., 2012), Mexico (Hebbeln et al., 2014), and Morocco (Foubert et al., 2008). The clustering of mounds is an issue open to discussion and various possible reasons triggering such clustering have been suggested. These range from antecedent structural topography and fault systems to regional productivity patterns and current systems (Wilson, 1975; Ahr, 1989; Brachert et al., 1992; Hovland et al., 1994; Bridges et al., 1995; Lees and Miller, 1995; Meyer et al., 1995; Bourque et al., 1995; Jeffery,

1997; Bourrouilh et al., 1998; Belka, 1998; Huvenne et al., 2007; Eisele et al., 2011; Hebbeln et al.,

2014). Thus, clustering of carbonate mounds also is a feature that is not linked to either shallow or ACCEPTED MANUSCRIPT bathyal settings.

5. Conclusions

Following some general ideas such as (1) that high carbonate production needed to create carbonate mounds has to involve photoautotrophic symbionts (Stanton, 2006; Pomar and Hallock, 2008), (2) that sedimentological indications for high energy levels reflect deposition near the fair-weather or storm wave base (Brachert et al., 1992; Tosolini et al., 2012), (3) that the occurrence of remains of photosynthetic, benthic organisms proves a deposition within the photic zone (e.g., Mamet and

18

ACCEPTED MANUSCRIPT

Preat, 2013), and some other indications as outlined above, the depositional environment of

Palaeozoic and Mesozoic carbonate mounds often has been linked to shallow shelf/ramp settings.

However, modern analogues for such shallow carbonate mounds are rare and do not match the wide distribution of fossil carbonate mounds.

All the commonly used criteria to allocate a shelf origin of ancient carbonate mounds, except the presence of photoautotrophic organisms, are also applicable for carbonate mounds being presently formed well below the shelf break down to water depths of ~1000 m along many continental margins in the world. This refers specifically to vigorous hydrodynamics that are even critical for the development of these modern bathyal mounds. Although the presence of benthic photoautotrophic organisms would be a clear indicator for shallow-water settings, for most of the organisms found in fossil carbonate mounds and being interpreted to be photoautotrophic, the taxonomic position and the metabolism are controversially discussed.

Interestingly, nowadays bathyal mounds appear to be much more common than shallow-water carbonate mounds. Considering the work done on modern carbonate mounds over the last two decades, it appears that for many ancient carbonate mounds a formation beyond the shelf break appears as likely, if not even more likely, as a formation in shallow shelf waters. Thus, as long as no proven photoautotrophic in-situ fossils are found, for many ancient carbonate mounds a reconsideration of the arguments used to conclude their depth of formation might be timely, ACCEPTED MANUSCRIPT especially as such conclusions can have large implications in terms of interpreting the palaeogeography and delineating the extension of former shelf seas. In addition, as modern analogues, recent carbonate mounds might serve as laboratories to gain a better process understanding to support the interpretation of ancient carbonate mounds. By comparing the substantial new knowledge about the distribution of modern carbonate mounds and about their controlling factors gained over the last ~20 years with the long history of research done on ancient carbonate mounds, this paper hopefully will stimulate an intense, open and fruitful dialogue

19

ACCEPTED MANUSCRIPT

between scientists studying either ancient or modern mound systems in order to advance our knowledge of carbonate mounds in general.

Acknowledgements

We thank the numerous colleagues who guided us in the field or supplied material for comparisons used in the paper, in particular the organizers of the COCARDE fieldtrip 2011 in Morocco that actually triggered the writing of this manuscript. We acknowledge funding by the Deutsche

Forschungsgemeinschaft (D.H., grant He 3412/17 - WACOM), the Swiss National Science Foundation

(E.S.) and the European Science Foundation in the frame of COCARDE. We are grateful to Gregor

Eberli, John Reijmer, André Strasser and two anonymous journal referees for commenting earlier versions of the manuscript.

References

Ahr, W.M., 1989. Sedimentary and tectonic controls on the development of an Early Mississippian

carbonate ramp, Sacramento Mountains area, New Mexico, in: Crevello, P.D., Wilson, J.L.,

Sarg, J.F., Read, J.F. (eds.), Controls on and Basin Development. SEPM

Spec. Publ. 44, 203-212.

Ahr, W.M., Stanton, R.J., 1994. Comparative sedimentology and paleontology of Waulsortian mounds ACCEPTED MANUSCRIPT and coeval level-bottom sediments of the Lower Lake Valley Formation (Lower Mississippian) in the Sacramento Mountains (New Mexico, USA). Abhandl. Geol.

Bundesanstalt 50, 11-24.

Aretz, M., Chevalier, E., 2007. After the collapse of stromatoporid-coral reefs-the Famennian and

Dinantian reefs of Belgium: much more than Waulsortian mounds. Spec. Publ. Geol. Soc.

London 275, 7-20.

Aretz, M., Herbig, H.-G., 2003. Coral-Rich Bioconstructions in the Visean (Late Mississippian) of

Southern Wales (Gower Peninsula, UK). Facies 49, 221-242.

20

ACCEPTED MANUSCRIPT

Bathurst, R.G.C., 1971. Carbonate Sediments and Their Diagenesis. Devel. Sediment. 12, 620 pp.

Belka, Z., 1981. The alleged algal genus Aphralysia is a formainifer. N. Jb. Geol. Paläo. Monatshefte

1981(5), 501-522.

Belka, Z., 1998. Early Devonian Kess-Kess carbonate mud mounds of the eastern Anti-Atlas

(Morocco), and their relation to submarine hydrothermal venting. J. Sedim. Res. 68, 368-

377.

Berra, F., Balini, M., Levera, M., Nicora, A., Salamati, R., 2012. Anatomy of carbonate mounds from

the Middle Anisian of Nakhlak (Central Iran): architecture and age of a subtidal microbial-

bioclastic carbonate factory. Facies 58, 685–705, doi:10.1007/s10347-012-0299-z.

Beuck, L., Freiwald, A., Taviani, M., 2010. Spatiotemporal bioerosion patterns in deep-water

scleractinians from off Santa Maria di Leuca (Apulia, Ionian Sea). Deep-Sea Res. II 57, 458-

470.

Bosence, D.W.J., 1995. Anatomy of a recent biodetrital mud-mound, Florida Bay, U.S.A., in: Monty,

C.L.V., Bosence, D.W.J., Bridges, P.H., Pratt, B.R. (eds.), Carbonate Mud-Mounds: Their

Origin and Evolution. IAS Spec. Publ. 23, pp. 475–493.

Bosence, D.W.J., Bridges, P.H., 1995. A review of the origin and evolution of carbonate mud-mounds,

in: Monty, C.L.V., Bosence, D.W.J., Bridges, P.H., Pratt, B.R. (eds.), Carbonate Mud-Mounds:

Their Origin and Evolution. IAS Spec. Publ. 23, pp. 3–10. ACCEPTED MANUSCRIPT Boulvain, F., 2001. Facies architecture and diagenesis of Belgian late Frasnian carbonate mounds (Petit-Mont Member). Sedim. Geol. 145, 269–294.

Boulvain, F., 2007. Frasnian carbonate mounds from Belgium: sedimentology and

palaeoceanography. Spec. Publ. Geol. Soc. London 275, 125–142.

Boulvain, F., Cornet, P., Da Silva, A.-C., Delaite, G., Demany, B., Humblet, M., Renard, M., Coen-

Aubert, M., 2004. Reconstructing -like mounds from the Frasnian of Belgium. Facies 50,

313–326.

21

ACCEPTED MANUSCRIPT

Bourque, P.-A., Boulvain, F., 1993. A model for the origin and petrogenesis of the red stromatactis

limestone of Palaeozoic carbonate mounds. J. Sedim. Petrol. 63, 607–619.

Bourque, P.A., Gignac, H., 1983. Sponge-constructed stromatactis mud mounds, Silurian of Gaspé,

Québec. J. Sedim. Petrol., 53, 521-532.

Bourque, P.A., Gignac, H., 1986. Sponge-constructed Stromatactis mud mounds, Silurian of Gaspe,

Quebec - Reply. J. Sedim. Petrol., 56, 461-463.

Bourque, P.-A., Madi, A., Mamet, B.L., 1995. Waulsortian-type bioherm development and response

to sea-level fluctuations: Upper Visean of Béchar Basin, western Algeria. J. Sedim. Res. B65,

80-95.

Bourrouilh, R., Bourque, P.-A., Dansereau, P., Bourrouilh-Le Jan, F., Weyant, P., 1998. Synsedimentary

tectonics, mud-mounds and sea-level changes on a Palaeozoic carbonate platform margin:

a Devonian Montagne Noire example (France). Sedim. Geol. 118, 95–118.

Brachert, T.C., 1991. Environmental control on fossilization of siliceous sponge assemblages: a

proposal, in: Reitner, J., Keupp, H. (eds.), Fossil and Recent Sponges. Springer, New York,

pp. 543-553.

Brachert, T.C., Buggisch, W., Flügel, E., Hüssner, H.M., Joachimski, M.M., Tourneur, F., Walliser, O.H.,

1992. Controls of mud mound formation: the Early Devonian Kess-Kess carbonates of the

Hamar Laghdad, Antiatlas, Morocco. Geol. Rundsch. 81, 15-44. ACCEPTED MANUSCRIPT Bridges, P.H., Chapman, A.J., 1988. The anatomy of a deep water mud-mound complex to the southwest of the Dinantian platform in Derbyshire, UK. Sedimentology 35, 139-162.

Bridges, P.H., Gutteridge, P., Pickard, N.A.H., 1995. The environmental setting of early carboniferous

mud-mounds, in: Monty, C.L.V., Bosence, D.W.J., Bridges, P.H., Pratt, B.R. (eds.), Carbonate

Mud-Mounds: Their Origin and Evolution. IAS Spec. Publ. 23, pp. 171–190.

Cairns, S.D., Stanley, G.D., Jr., 1981. Ahermatypic coral banks: living and fossil counterparts.

Proceedings Fourth Int. Coral Reef Symposium, Manila, v. 1, p. 611-618.

22

ACCEPTED MANUSCRIPT

Colman, J.G, Gordon, D.M, Lane, A.P., Forde, M.J., Fitzpatrick, J.J., 2005. Carbonate mounds off

Mauretania, Northwest Africa: status of deep-water corals and implications for

management of fishing and oil exploration activities, in: Freiwald, A., Roberts, J.M. (eds.),

Cold-water Corals and Ecosystems. Springer-Verlag, Berlin Heidelberg, pp. 417-441

Conway, K.W., Krautter, M., Barrie, J.V., Whitney, F., Thomson, R.E., Reiswig, H., Lehnert, H., Mungov,

G., Bertram, M., 2005. Sponge reefs in the Queen Charlotte Basin, Canada: controls on

distribution, growth and development, in: Freiwald, A., Roberts, J.M. (eds.), Cold-Water

Corals and Ecosystems. Springer-Verlag, Berlin Heidelberg, pp. 605-621.

Correa, T., Grasmück, M., Eberli, G.P., Reed, J.K., Verwer, K., Purkis, S., 2012. Variability of cold-water

coral mounds in a high sediment input and tidal current regime, Straits of Florida.

Sedimentology 59, 1278–1304.

De Mol, B., Van Rensbergen, P., Pillen, S., Van Herreweghe, K., Van Rooij, D., McDonnell, A.,

Huvenne, V., Ivanov, M., Swennen, R., Henriet, J.P., 2002. Large deep-water coral banks in

the Porcupine Basin, southwest of Ireland. Mar. Geol. 188, 193–231.

Dorschel, B., Hebbeln D., Rüggeberg, A., Dullo, W.-Ch., Freiwald, A., 2005. Growth and erosion of a

cold-water coral covered carbonate mound in the Northeast Atlantic during the Late

Pleistocene and Holocene. Earth Planet. Sci. Let. 233, 33–44.

Dorschel, B., Hebbeln D., Rüggeberg, A., Dullo, W.-Ch., 2007. Carbonate budget of a cold-water coral ACCEPTED MANUSCRIPT carbonate mound: Propeller Mound, Porcupine Seabight. Int. J. Earth Sci. 96, 73–83. Dorschel, B., Wheeler, A.J., Monteys, X., Verbruggen, K., 2010. Atlas of the Deep-Water :

Ireland. Springer, Dordrecht Heidelberg London New York. 164 p., doi:10.1007/978-90-481-

9376-9.

Dullo, W.-C., 1990. Facies, fossil record and age of Pleistocene reefs from the Red Sea (Saudi Arabia).

Facies 22, 1–46.

Eisele, M., Hebbeln, D., Wienberg, C., 2008. Growth history of a cold-water coral covered carbonate

mound — Galway Mound, Porcupine Seabight, NE-Atlantic. Mar. Geol. 253, 160-169.

23

ACCEPTED MANUSCRIPT

Eisele, M., Frank, N., Wienberg, C., Hebbeln, D., López Correa, M., Douville, E., Freiwald A., 2011.

Productivity controlled cold-water coral growth periods during the last glacial off

Mauritania. Mar. Geol. 280, 143-149.

Ferdelman, T.G., Kano, A., Williams, T., Henriet, J.-P., IODP Expedition 307 Scientists, 2006. Modern

carbonate mounds: Porcupine Drilling, in Proc. Integrated Ocean Drilling Program 307,

College Station, Texas, Integrated Ocean Drilling Program Management International, Inc.,

doi:10.2204/iodp.proc.307.2006.

Fink, H.G., Wienberg, C., Hebbeln, D., McGregor, H.V., Schmiedl, G., Taviani, M., Freiwald, A., 2012.

Oxygen control on Holocene cold-water coral development in the eastern Mediterranean

Sea. Deep-Sea Res. I 62, 89–96.

Fink, H.G., Wienberg, C., Pol-Holtz, R., Wintersteller, P., Hebbeln, D., 2013. Cold-water coral growth in

the Alboran Sea related to high productivity during the Late Pleistocene and Holocene. Mar.

Geol. 339, 71–82.

Flügel, E., 2004. Microfacies of Carbonate Rocks: Analysis, Interpretation and Application. Springer,

Berlin, 976 pp.

l gel, E., 2010. Microfacies of Carbonate Rocks: Analysis, Interpretation and Application, 2nd ed.

Springer, Berlin, 984 pp.Flügel, E., Flügel-Kahler, E., 1992. Phanerozoic reef evolution: Basics

questions and data base. Facies 26, 167-278. ACCEPTED MANUSCRIPT Fosså, J., Lindberg, B., Christensen, O., Lundälv, T., Svellingen, I., Mortensen, P., Alvsvåg, J., 2005. Mapping of Lophelia reefs in Norway: experiences and survey methods, in: Freiwald, A.,

Roberts, J.M. (eds.), Cold-Water Corals and Ecosystems. Springer Berlin Heidelberg, pp. 359-

391.

Foubert, A., Depreiter, D., Beck, T., Maignien, L., Pannemans, B., Frank, N., Blamart, D., Henriet, J.P.,

2008. Carbonate mounds in a mud volcano province off northwest Morocco: key to

processes and controls. Mar. Geol. 248, 74-96.

24

ACCEPTED MANUSCRIPT

Frank, N., Freiwald, A., López Correa, M., Wienberg, C., Eisele, M., Hebbeln, D., Van Rooij, D., Henriet,

J.P., Colin, C., van Weering, T., de Haas, H., Buhl-Mortensen, P., Roberts, J.M., De Mol, B.,

Douville, E., Blamart, D., Hatte, C., 2011. Northeastern Atlantic cold-water coral reefs and

climate. Geol. 39, 743–746.

Freiwald, A., Henrich, R., Pätzold, J., 1997. Anatomy of a deep-water coral reef from Stjernsund, West

Finmark, Northern Norway, in: James, N.P., Clarke, J.A.D. (eds.), Cool-Water Carbonates,

SEPM Spec. Publ. 56, 141-162.

Freiwald, A., Beuck, L., Rüggeberg, A., Taviani, M., Hebbeln, D., R/V Meteor M70-1 participants, 2009.

The white coral community in the Central Mediterranean Sea - revealed by ROV surveys.

Oceanography 22, 36-52.

Geldsetzer, H.H.J., James, N.P., Tebbutt, G.E. (eds.), 1989. Reefs, Canada and Adjacent Areas. Can.

Soc. Petrol. Geol. Mem. 13, 775 pp.

Grasmueck, M., Eberli, G.P., Viggiano, D.A., Correa, T., Rathwell, G., Luo, J., 2006. Autonomous

underwater vehicle (AUV) mapping reveals coral mound distribution, morphology, and

in deep water of the Straits of Florida. Geophys. Res. Let. 33, L23616,

doi:10.1029/2006GL027734.

Grasmueck, M., Eberli, G.P., Correa, T., Viggiano, D. A., Luo, J., Wyatt, G.J., Wright, A.E., Reed, J.,

2007. AUV-based environmental characterization of deep-water coral mounds in the Straits ACCEPTED MANUSCRIPT of Florida. OTC, 18510, 1–11. Gutteridge, P., 1995. Late Dinantian (Brigantian) carbonate mud-mounds of the Derbyshire carbonate

platform, in: Monty, C.L., Bosence, D.W.J., Bridges, P.H. (eds.), Carbonate Mud-Mounds:

Their Origin and Evolution. IAS Spec. Publ. 23, pp. 289-307.

Hebbeln, D., Wienberg, C., Wintersteller, P., Freiwald, A., Becker, M., Beuck, L., Dullo, C., Eberli, G.P.,

Glogowski, S., Matos, L., Forster, N., Reyes-Bonilla, H., Taviani, M., MSM 20-4 shipboard

scientific party, 2014. Environmental forcing of the Campeche cold-water coral province,

southern Gulf of Mexico. Biogeosci. 11, 1799-1815, doi:10.5194/bg-11-1799-2014.

25

ACCEPTED MANUSCRIPT

Henriet, J.P., Hamoumi, N., Da Silva, A.C., Foubert, A., Lauridsen, B.W., Rüggeberg, A., Van Rooij, D.,

2014. Carbonate mounds: from paradox to World Heritage, Mar. Geol. 352, 89-110,

doi:10.1016/j.margeo.2014.01.008.

Hook, J.E., Golubic, S., Milliman, J.D., 1984. Micritic cement in microborings is not necessarily a

shallow-water indicator. J. Sedim. Petrol. 54, 425-431.

Hovland, M., Croker, P.F., Martin, M., 1994. Fault-associated seabed mounds (carbonate knolls?) off

western Ireland and north-west Australia, Mar. Pet. Geol. 11, 232– 246.

Huvenne, V.A.I., Beyer, A., de Haas, H., Dekindt, K., Henriet, J.P., Kozachenko, M., Olu-Le Roy, K.,

Wheeler, A.J., the TOBI/Pelagia 197, CARACOLE cruise participants, 2005. The seabed

appearance of different coral bank provinces in the Porcupine Seabight, NE Atlantic: results

from sidescan sonar and ROV seabed mapping, in: Freiwald, A., Roberts, J.M. (eds.), Cold-

Water Corals and Ecosystems. Springer-Verlag, Berlin Heidelberg, pp. 535-569.

Huvenne, V.A.I., Bailey, W.R., Shannon. P.M., Naeth, J., Di Primio, R., Henriet, J.P., Horsfield, B., De

Haas, H., Wheeler, A, Olu, K., 2007. The Magellan mound province in the Porcupine Basin. .

Int. J. Earth Sci. 96, 85-101.

James, N.P., Bourque, P.-A., 1992. Reefs and mounds, in: Walker, R.G., James, N.P. (eds.), Facies

Models: Response to Change. Geological Association of Canada, St. John's,

Newfoundland, pp. 323-347. ACCEPTED MANUSCRIPT James, N.P., Wood, R.A., 2010. Reefs and Reef Mounds, in: James, N.P., Walker, N. (eds.), Facies Models 4. Geological Association of Canada. Series GEOtext, pp. 421-447.

James, N., Feary, D.A., Surlyk, F., Simo, J.A., Betzler, C., Holbourn, A.E., Li, Q., Matsuda, H.,

Machiyama, H., Brooks, G., Andres, M.S., Hine, A.C., Malone, M.J., Ocean Drilling Program

Leg 182 Scientific Party, 2000. bryozoan mounds in cool-water, upper slope

environments: Great Australian Bight. Geol. 28, 647-650.

Jeffery, D.L., 1997. Tectonic control on sea-floor relief and the localization of Lower Mississippian

Waulsortian mounds, New Mexico. Geol. 25, 1011-1014.

26

ACCEPTED MANUSCRIPT

Jeffery, D.L., Stanton, R.J., 1996. Biotic gradients on a homoclinal ramp, the Alamogordo Member of

the Lake Valley Formation, Lower Misssissippian, New Mexico, USA, in: Strogen, P.,

Somerville, I.D., Jones, G.L., (eds.), Recent Advances in Lower Carboniferous Geology. Spec.

Publ. Geol. Soc. London 107, 111-126.

Kano, A., Ferdelman, T.G., Williams, T., Henriet, J.-P., Ishikawa, T., Kawagoe, N., Takashima, C.,

Kakizaki, Y., Abe, K., Sakai, S., Browning, E. L., Li, X., Integrated Ocean Drilling Program

Expedition 307 Scientists, 2007. Age constraints on the origin and growth history of a deep-

water coral mound in the northeast Atlantic drilled during Integrated Ocean Drilling

Program Expedition 307. Geol. 35, 1051-1054.

Kiessling, W., Flügel, E., Golonka, J. (eds.), 2002. Phanerozoic reef patterns. SEPM Spec. Publ. 72, 775

p.

Kirkby, K.C., Hunt, D., 1996. Episodic growth of a Waulsortian buildup: the Lower Carboniferous

Muleshoe Mound, Sacramento Mountains, New Mexico, USA, in: Strogen, P., Somerville,

I.D., Jones, G.L. (eds.), Recent Advances in Lower Carboniferous Geology. Spec. Publ. Geol.

Soc. London107, 97-110.

Lavaleye, M., Duineveld, G., Lundälv, T., White, M., Guihen, D., Kiriakoulakis, K., Wolff, G.A., 2009.

Cold-Water Corals on the Tisler Reef. Preliminary observations on the dynamic reef

environment. Oceanogr. 22, 76-84. ACCEPTED MANUSCRIPT Le Guilloux, E., Olu, K., Bourillet, J.-F., Savoye, B., Iglesias, S.P., Sibuet, M., 2009. First observations of deep-sea coral reefs along the Angola margin. Deep Sea Res. II 56, 2394-2403.

Lees, A., 1964. The structure and origin of the Waulsortian (Lower Carboniferous) 'reefs' of west-

central Eire. Phil. Trans. Royal Soc. London, Ser. B 247, 483-531.

Lees, A., Conil, R., 1980. The Waulsortian reefs of Belgium. Géobios Mem. Spécial 4, 35-46.

Lees, A., Miller, J., 1985. Facies variation in Waulsortian buildups, Part 2; Mid-Dinantian buildups

from Europe and North America. Geol. J. 20, 159-180.

27

ACCEPTED MANUSCRIPT

Lees, A., Miller, J., 1995. Waulsortian banks, in: Monty, C.L.V., Bosence, D.W.J., Bridges, P.H., Pratt,

B.R. (eds.), Carbonate Mud-Mounds: Their Origin and Evolution. IAS Spec. Publ. 23, pp.

191–272.

Lees, A., Hallet, V., Hibo, D., 1985. Facies variation in Waulsortian buildups, Part 1; A model from

Belgium. Geol. J. 20, 133-158.

Madi, A., Bourque, P.A., Mamet, B.L., 1996. Depth-related ecological zonation of a Carboniferous

carbonate ramp: Upper Viséan of Béchar Basin, Western Algeria. Facies 35, 58-80.

Mamet, B., 1991. Carboniferous calcareous algae, in: Riding, R. (ed.), Calcareous Algae and

Stromatolites. Springer-Verlag, New York, pp. 370-451.

Mamet, B., Preat, A., 2013. Essai de description d’algues nouvelles Paléozoïques. Geol. Belgica 16/1-

2, 35-48.

Meyer, D.L., Ausich, W.I., Bohl, D.T., Norris, W.A., Potter, P.E., 1995. Carbonate Mud-Mounds in the

Fort Payne Formation (Lower Carboniferous), Cumberland Saddle Region, Kentucky and

Tennessee, USA, in: Monty, C.L.V., Bosence, D.W.J., Bridges, P.H., Pratt, B.R. (eds.),

Carbonate Mud-Mounds: Their Origin and Evolution. IAS Spec. Publ. 23, pp. 273-287.

Mienis, F., De Stigter, H.C., De Haas, H., Van Weering, T.C.E., 2009a. Near-bed particle deposition and

resuspension in a cold-water coral mound area at the Southwest Rockall Trough margin, NE

Atlantic. Deep Sea Res. I 56, 1026-1038. ACCEPTED MANUSCRIPT Mienis, F., Van der Land, C., De Stigter, H.C., Van de Vorstenbosch, M., De Haas, H., Richter, T., Van Weering, T.C.E., 2009b. Sediment accumulation on a cold-water carbonate mound at the

southwest Rockall trough margin. Mar. Geol., 265 40-50.

Mienis, F., Duineveld, G.C.A., Davies, A.J., Ross, S.W., Seim, H., Bane, J., Van Weering, T.C.E., 2012.

The influence of nearbed hydrodynamic conditions on cold-water corals in the Viosca Knoll

area, Gulf of Mexico. Deep-Sea Res. I 60, 32–45.

Mienis, F., Duineveld, G.C.A., Davies, A.J., Lavaleye, M.M.S., Ross, S.W., Seim, H., Bane, J., van Haren,

H., Bergman, M.J.N., de Haas, H., Brooke, S., van Weering, T.C.E., 2014. Cold-water coral

28

ACCEPTED MANUSCRIPT

growth under extreme environmental conditions, the Cape Lookout area, NW Atlantic,

Biogeosci. 11, 2543-2560, doi:10.5194/bg-11-2543-2014.

Monty, C.L.V., Bosence, D.W.J., Bridges, P.H., Pratt, B.R. (eds.), 1995. Carbonate Mud-Mounds: Their

Origin and Evolution. IAS Spec. Publ. 23, 537pp.

Monty, C.L.V., 1995. The rise and nature of carbonate mud-mounds: an introductory actualistic

approach, in: Monty, C.L.V., Bosence, D.W.J., Bridges, P.H., Pratt, B.R. (eds.), Carbonate

Mud-Mounds: Their Origin and Evolution. IAS Spec. Publ. 23, pp. 11–48.

Mullins, H.T., Newton, C.R., Heath, K., Van Buren, H.M., 1981. Modern deep-water coral mounds

north of Little Bahama Bank: criteria for the recognition of deep-water coral bioherms in

the rock record. J. Sedim. Petrol. 51, 999-1013.

Murphy, R.X., 1988. Facies variations within the Waulsortian Limestone Formation of Dunvergan

Syncline, southern Ireland. Proc. Geol. Ass. 99, 205-219.

Nose M., Schmid, D.U., Leinfelder, R.R., 2006. Significance of microbialites, calcimicrobes, and

calcareous algae in reefal framework formation from the Silurian of Gotland, Sweden.

Sedim. Geol. 192, 243-265.

Paull, C.K., Neumann, A.C., am Ende, B.A., Ussler III, W., Rodriguez, N.M., 2000. Lithoherms on the

Florida–Hatteras slope. Mar. Geol. 166, 83–101

Pomar, L., Hallock, P., 2008. Carbonate factories: a conundrum in sedimentary geology. Earth-Sci. ACCEPTED MANUSCRIPT Rev. 87, 134–169. Pratt, B.R., 1995. The origin, biota and evolution of deep-water mud-mounds. In: Monty, C.L.V.,

Bosence, D.W.J., Bridges, P.H., Pratt, B.R. (eds.), Carbonate Mud-Mounds: Their Origin and

Evolution. IAS Spec. Publ. 23, pp. 49–126.

Pray, L.C., 1958. Fenestrate bryozoan core facies, Mississippian bioherms, southwestern United

States. J. Sedim. Petrol. 28, 261-273.

Remia, A., Taviani, M., 2005. Shallow-buried Pleistocene Madrepora-dominated coral mounds on a

muddy continental slope, Tuscan Archipelago, NE Tyrrhenian Sea. Facies 50, 419–425.

29

ACCEPTED MANUSCRIPT

Reveillaud, J., Freiwald, A., Van Rooij, D., Le Guilloux, E., Altuna, A., Foubert, A., Vanreusel, A., Olu-

LeRoy, K., Henriet, J. P., 2008. The distribution of scleractinian corals in the Bay of Biscay,

NE Atlantic. Facies 54, 317-331.

Riding, R., 1991. Classification of microbial carbonates, in: R. Riding (ed.), Calcareous Algae and

Stromatolites. Springer-Verlag, Berlin, pp. 21-51.

Reyes, J., Santodomingo, N., Gracia, A., Borrero-Perez, G., Navas, G., Mejia-Ladino, L.M., Bermudez,

A., Benavides, M., 2005. Southern Caribbean azooxanthellate coral communities off

Colombia, in: Freiwald, A., Roberts, J.M. (eds.), Cold-Water Corals and Ecosystems.

Springer-Verlag, Berlin Heidelberg, pp. 309-330.

Riding, R., 2002. Structure and composition of organic reefs and carbonate mud mounds: concepts

and categories. Earth-Sci. Rev. 58, 163–231.

Riding, R., Soja, C.M., 1993. Silurian calcareous algae, cyanobacteria, and microproblematica from the

Alexander Terrane, Alaska. J. Palaeont. 67, 710-728.

Roberts J.M., Wheeler, A.J., Freiwald, A., 2006. Reefs of the deep: the biology and geology of bold-

water coral ecosystems. Science 312, 543-547.

Roberts, J.M., Wheeler, A.J., Freiwald, A., Cairns, S.D., 2009. Cold-Water Corals: The Biology and

Geology of Deep-Sea Coral Habitats. Cambridge University Press, Cambridge, 334 p.

Samankassou, E., Allmen, V.K., Bahamonde, J.R., 2013. Growth dynamics of Pennsylvanian carbonate ACCEPTED MANUSCRIPT mounds from a mixed terrigenous-carbonate ramp In the Puebla de Lillo area, Cantabrian Mountains, Northern Spain. J. Sedime. Res. 83, 1099-1112.

Schlager, W., 2003. Benthic carbonate factories of the Phanerozoic. Int. J. Earth Sci. 92, 445-464.

Schmid, D.U., Leinfelder, R.R., Nose, M., 2001. Growth dynamics and ecology of Upper Jurassic

mounds, with comparisons to Mid-Palaeozoic mounds. Sedim. Geol. 145, 343– 376.

Seibold, E., Berger, W.H., 1993. The Sea Floor. Springer, Berlin, Heidelberg, 356 p.

Stanton, R.J. Jr., 2006. Nutrient models for the development and location of ancient reefs. Geo. Alp 3,

191-206.

30

ACCEPTED MANUSCRIPT

Stanton, R.J. Jr., Jeffery, D.L., Guillemette, R.N., 2000. Oxygen minimum zone and internal waves as

potential controls on location and growth of Waulsortian mounds (Mississippian,

Sacramento Mountains, New Mexico). Facies 42, 161-176.

Taviani, M., Remia, A., Corselli, C., Freiwald, A., Malinverno, E., Mastrototaro, F., Savini, A., Tursi, A.,

2005a. First geo-marine survey of living cold-water Lophelia reefs in the Ionian Sea

(Mediterranean basin). Facies 50, 409–417.

Strogen, P., Somerville, I.D., Jones, G.LL. (eds.), 1996. Recent Advances in Lower Carboniferous

Geology. Spec. Publ. Geol. Soc. London 107, 448 pp.

Titschack, J., Thierens, M., Dorschel, B., Schulbert, C., Freiwald, A., Kano, A., Takashima, C., Kawagoe,

N., Li, X., IODP Expedition 307 scientific party, 2009. Carbonate Budget of a cold-water coral

mound (Challenger Mound, IODP Exp. 307). Mar. Geol. 259, 36-46.

Tsien, H.-H., Mouravieff, A.N., Mountjoy, E.W., 1980. Devonian reefs in Belgium. Geobios – Mem.

Spec. 4, 17–33.

Tosolini, A.-M.P., Wallace, W.M., Gallagher, S.J., 2012. Shallow water mud-mounds of the Early

Devonian Buchan Group, East Gippsland, Australia. Sedim. Geol. 281, 208-221.

Vachard, D., Aretz, M., 2004. Biostratigraphical precisions on the Early Serpukhovian (Late

Mississippian), by means of a carbonate algal microflora (cyanobacteria, algae and pseudo-

algae) from La Serre (Montagne Noire, France). Geobios 37(5), 643-666 ACCEPTED MANUSCRIPT Vachard, D., Sommerville, I.D., Cozar, P., 2004. Fasciella and Praedonezella (Mississippian-early Pennsylvanian algae): revision and new species. Rev. Española Micropaleont. 36(2), 263-

278.

Van der Land, C., Eisele, M., Mienis, F., de Haas, H., Hebbeln, D., Reijmer, J.J.G., van Weering, T.C.E.,

2014. Carbonate mound development in contrasting settings on the Irish margin. Deep Sea

Res. II 99, 297-306.

31

ACCEPTED MANUSCRIPT

Viana, A.R., Faugres, J.C., Kowsmann, R.O., Lima, J.A.M., Caddah, L.E.G., Rizzo, J.G., 1998. Hydrology,

morphology and sedimentology of the Campos continental margin, offshore Brazil. Sedim.

Geol. 115, 133-157.

Walther, J., 1894. Einleitung in die Geologie als historische Wissenschaft. Verlag von Gustav Fischer,

Jena, 3 volumes., 1055 pp.

Webb, G.E., 1996. Was Phanerozoic reef history controlled by the distribution of non-enzymatically

secreted reef carbonates (microbial carbonate and biologically induced cement)?

Sedimentol. 43, 947-971.

Webb, G.E., 2001. Famennian mud-mounds in the proximal fore-reef slope, Canning Basin, Western

Australia. Sedim. Geol. 145, 295– 315.

Webb, G.E., 2002. Late Devonian and Early Carboniferous reefs: depressed reef building following the

middle Paleozoic collapse. In: Flügel, E., Kiessling, W., Golonka, J. (eds.) Phanerozoic Reef

Patterns. SEPM Spec. Publ. 72, 239-269.

Wendt, J., 1993. Steep-sided carbonate mud-mounds in the Middle Devonian of the eastern Anti-

Atlas, Morocco. Geol. Magaz. 130, 69–83.

Wendt, J., Belka, Z., Moussine-Pouchkine, A., 1993. New architectures of deep-water carbonate

buildups: evolution of mud-mounds into mud ridges (Middle Devonian, Algerian Sahara).

Geol. 21, 723-726. ACCEPTED MANUSCRIPT Wendt, J., Belka, Z., Kaufmann, B., Kostrewa, R., Hayer, J., 1997. The world's most spectacular mud mounds (Middle Devonian, Algerian Sahara). J. Sedim. Res. 67, 424-436.

Wendt, J., Kaufmann, B., Belka, Z., 2001. An exhumed Palaeozoic underwater scenery: the Visean

mud mounds of the eastern Anti-Atlas (Morocco). Sedim. Geol. 145, 215– 233.

Wheeler, A.J.P, Beyer, A., Freiwald, A., de Haas, H., Huvenne, V.A.I., Kozachenko, M., Olu-Le Roy, K.,

Opderbecke, J., 2007. Morphology and environment of cold-water coral carbonate mounds

on the NW European margin. Int. J. Earth Sci. 96, 37–56, doi:10.1007/s00531-006-0130-6

32

ACCEPTED MANUSCRIPT

White, M., Dorschel, B., 2010. The importance of the permanent to the cold water coral

carbonate mound distribution in the NE Atlantic. Earth Planet. Sci. Let. 296, 395–402.

Wienberg, C., Beuck, L., Heidkamp, S., Hebbeln, D., Freiwald, A., Pfannkuche, O., Monteys, X., 2008.

Franken Mound – facies and biocoenoses on a newly-discovered 'carbonate mound' on the

western Rockall Bank, NE Atlantic. Facies 54, 1-24, doi:10.1007/s10347-007-0118-0.

Wienberg, C., Frank, N., Mertens, K., Stuut, J.-B., Marchant, M., Fietzke, J., Mienis, F., Hebbeln, D.,

2010. Glacial cold-water corals growth in the Gulf of Cádiz: implications of increased

palaeo-productivity. Earth Planet. Sci. Let. 298, 405-416, doi:10.1016/j.epsl.2010.08.017

Wilber, R.J., Neumann, A.C., 1993. Effects of Submarine Cementation on Microfabrics and Physical

Properties of Carbonate Slope Deposits, Northern Bahamas, in: Rezak, R., Lavoie, D.L. (eds.),

Carbonate Microfabrics. Frontiers in Sedimentary Geology, Springer-Verlag New York, pp.

79-94.

Wilson, J.L., 1975. Carbonate Facies in Geologic History. Springer-Verlag, Berlin, pp. 1–471.

Wood, A., 1948. Sphaerocodium, a misinterpreted fossil from the Wenlock Limestone. Proc. Geol.

Ass. 59, 9-22.

Wood, R., 1998. The ecological evolution of reefs. Ann. Rev. Ecol. Systematics 29, 179–206.

Wood, R., 2001. Are reefs and mud-mounds really so different? Sedim. Geol. 145, 161–171.

Wray, J.L., 1972. Environmental distribution of calcareous algae in Upper Devonian reef complexes. ACCEPTED MANUSCRIPT Geol. Rundsch. 61, 578–584.

33

ACCEPTED MANUSCRIPT

Figure captions

For the printed version of the manuscript the use of black & white figures is intended, while colour figures should be used for the online version.

Fig. 1: Distribution of cold-water corals (dots and lines in light blue, after Roberts et al., 2006) in the

Atlantic Ocean. Superimposed are examples of proven occurrences of recent (stars, mostly representing clusters of mounds - compiled - compiled from Colman et al., 2005; Fink et al., 2013;

Foubert et al., 2008, Grasmueck et al., 2006; Hebbeln et al., 2014; Le Guilloux et al., 2009, Mienis et al. 2012; 2014; Mullins et al., 1981; Remia and Taviani, 2005; Reveillaud et al., 2008; Reyes et al.,

2005; Taviani et al., 2005; Viana et al., 1998; Wheeler et al. 2007) and of ancient carbonate mounds

(triangles, compiled from compiled from Wilson, 1975; Geldsetzer et al., 1989; Flügel and Flügel-

Kahler, 1992; Bourque et al., 1995; Pratt, 1995; Gutteridge et al., 1995; Lees and Miller, 1995; Madi et al., 1996; Strogen et al., 1996; Boulvain, 2001; Wendt et al., 2001, Riding, 2002; Webb, 2002). At many continental margin sites for which the occurrence of cold-water corals has been described, modern carbonate mounds still might be discovered, as recently e.g. along the Angolan margin (Le

Guilloux et al., 2009), the Moroccan Mediterranean margin (Fink et al., 2013), and the Mexican

Campeche Bank (Hebbeln et al., 2014). ACCEPTED MANUSCRIPT

Fig. 2: Examples for the shape and dimensions of ancient and modern carbonate mounds. Upper panel: Province of exhumed mounds (Kess Kess mounds, Devonian) from the central area of Hamar

Laghdad Ridge, Anti-Atlas, Morocco (details on composition of these mounds are provided in

Brachert et al., 1992). Individual mounds on the image are in average 30 m high. Lower panel:

Multibeam echosounder map of cold water coral carbonate mounds at the Moroccan Atlantic continental margin with dimensions indicated. Data have been obtained in 2014 during expedition

MSM-36 with the German RV Maria S. Merian (Hebbeln et al., unpubl. data).

34

ACCEPTED MANUSCRIPT

Fig. 3: Comparison of the lithology of ancient and modern carbonate mounds. (a) Vertical cut-surface of sediment core GeoB 14904-2 from a Mauritanian margin cold-water coral mound (to the left) compared to a vertical outcrop surface (to the right) of a Visean carbonate mound, Moroccan Anti-

Atlas (for the setting see Wendt et al., 2001). Scale bar is 15 cm. The white rectangle indicates the zoom-in displayed in (b).

Fig. 4: Rippled sea floor with coral fragments from the Porcupine Seabight at the Irish margin in a water depth of ~600 m. The two dots are 20 cm apart from each other. The photograph has been taken in 2010 during expedition POS400 with the German RV Poseidon with the ROV CHEROKEE

(MARUM, University of Bremen, Germany).

Fig. 5: Development from the living coral thicket towards the sedimentary record. (a) Living Lopehlia pertusa framework (in pink) growing on dead, but still intact framework (grey to white) in ~640 m water depth at Rockall Bank, NE Atlantic. The entire thicket reaches ca. 1 m in height. The photograph has been taken in 2004 during expedition M61-3 with the German RV Meteor with the

ROV QUEST (MARUM, University of Bremen, Germany). (b) Coral rubble facies in the Porcupine

Seabight at the Irish margin in ~750 m water depth (width of the image: ~ 30 cm). The photograph ACCEPTED MANUSCRIPT has been taken in 2010 during expedition POS400 with the German RV Poseidon with the ROV CHEROKEE (MARUM, University of Bremen, Germany). (c) Computer-tomograph (CT) image of sediment core GeoB 11569-2 collected from a Mauritanian margin cold-water coral mound (width:

12 cm) showing that generally disintegrated coral fragments (visualized here by CT technique without surrounding sediment) form the mound facies.

35

ACCEPTED MANUSCRIPT

Fig. 6: Detailed bathymetric 3D-map for Propeller Mound, an actively growing cold-water coral carbonate mound in ~800 m water depth at the Irish continental margin (Huvenne et al., 2005). The map clearly shows the flattened top of the mound (map courtesy by B. Dorschel).

Figure 7

(a) Sketch of a Middle Devonian sequence at Jebel el Oftal, Morocco, including carbonate mounds and the surrounding bedded limestone. Redrawn from Wendt (1993).

(b) Sequence of Devonian Kess Kess mounds and the surrounding bedded limestone of Hamar

Laghdad Ridge, Anti-Atlas, Morocco. Mounds seem to be amalgamated and/or aligned. As in the examples reported by Wendt (1993), the facies architecture lacks vertically superimposed facies

(mounds versus bedded limestone) indicative of prograding and/or shallowing-upwards sequences.

The mounds reach ca. 30 m in height.

ACCEPTED MANUSCRIPT

36

ACCEPTED MANUSCRIPT

Fig 1

ACCEPTED MANUSCRIPT

37

ACCEPTED MANUSCRIPT

Fig 2

ACCEPTED MANUSCRIPT

38

ACCEPTED MANUSCRIPT

ACCEPTED MANUSCRIPT

Fig 3

39

ACCEPTED MANUSCRIPT

Fig 4

ACCEPTED MANUSCRIPT

40

ACCEPTED MANUSCRIPT

Fig 5

ACCEPTED MANUSCRIPT

41

ACCEPTED MANUSCRIPT

Fig 6

ACCEPTED MANUSCRIPT

42

ACCEPTED MANUSCRIPT

Fig 7

ACCEPTED MANUSCRIPT

43