background image

GEOLOGICA CARPATHICA

, DECEMBER 2018, 69, 6, 573–592

doi: 10.1515/geoca-2018-0034

www.geologicacarpathica.com

Palaeoenvironmental analysis of the Miocene barnacle 

facies: case studies from Europe and South America

GIOVANNI COLETTI

1,

, GIULIA BOSIO

1

, ALBERTO COLLARETA

2, 3

, JOHN BUCKERIDGE

4

,  

SIRIO CONSANI

5

 and AKRAM EL KATEB

6

1 

Dipartimento di Scienze dell’Ambiente e della Terra, Università di Milano-Bicocca, Piazza della Scienza 4, Milano 20126, Italy; 

 giovanni.p.m.coletti@gmail.com

2 

Dipartimento di Scienze della Terra, Università di Pisa, Via Santa Maria 53, Pisa 56126, Italy

3

Museo di Storia Naturale, Università di Pisa, via Roma 79, Calci 56011, Italy

4 

Earth & Oceanic Systems Group, RMIT, Melbourne 3001, Australia

5 

Dipartimento di Scienze della Terra, dell’Ambiente e della Vita, Università degli Studi di Genova, Corso Europa 26, Genova 16132, Italy

6 

Department of Geosciences, University of Fribourg, Chemin du Musée 6, Fribourg 1700, Switzerland

(Manuscript received August 1, 2018; accepted in revised form November 28, 2018)

Abstract: Acorn barnacles are sessile crustaceans common in shallow-water settings, both in modern oceans and in  

the Miocene geological record. Barnacle-rich facies occur from polar to equatorial latitudes, generally associated with 

shallow-water, high-energy, hard substrates. The aim of this work is to investigate this type of facies by analysing, from 

the palaeontological, sedimentological and petrographical points of view, early Miocene examples from Northern Italy, 

Southern France and South-western Peru. Our results are then compared with the existing information on both modern 

and fossil barnacle-rich deposits. The studied facies can be divided into two groups. The first one consists of very shallow, 

nearshore assemblages where barnacles are associated with an abundant hard-substrate biota (e.g., barnamol). The second 

one includes a barnacle-coralline algae association, here named “barnalgal” (= barnacle / red algal dominated), related to 

a deeper setting. The same pattern occurs in the distribution of both fossil and recent barnacle facies. The majority of them 

are related to very shallow, high-energy, hard-substrate, a setting that represents the environmental optimum for  

the development of barnacle facies, but exceptions do occur. These atypical facies can be identified through a complete 

analysis of both the skeletal assemblage and the barnacle association, showing that barnacle palaeontology can be  

a powerful tool for palaeoenvironmental reconstruction.

Keywords: Carbonate Factories, Heterozoan, Barnamol, Barnalgal, Tertiary Piedmont Basin, Sommières Basin,  

Pisco Basin.

Introduction

Acorn barnacles (Cirripedia: Sessilia) are common carbonate 

producers in modern and fossil shallow-water shelf environ-

ments (Foster 1987; Foster & Buckeridge 1987; Doyle et al. 

1997). Albeit often overlooked, this group of sessile, suspen-

sion-feeding crustaceans occur on any available surface in 

shallow seas, including “mobile surfaces” like turtles and 

whales (Ross & Newman 1967; Newman & Abbot 1980; 

Scarff 1986; Seilacher 2005; Bianucci et al. 2006; Dominici et 

al.  2011;  Harzhauser  et  al.  2011;  Collareta  et  al.  2016 a, b). 

They are common at middle and high latitudes (Raymond & 

Stetson 1932; Hoskin & Nelson 1969; Milliman 1972; Müller 

& Milliman 1973; Farrow et al. 1978; Hottinger 1983; Domack 

1988; Nelson et al. 1988; Scoffin 1988; Wilson 1988; Taviani 

et al. 1993; Henrich et al. 1995; Frank et al. 2014; Buckeridge 

2015), but they can also thrive at low-latitudes, especially  

in nutrient-rich environments (Glynn & Wellington 1983;  

Carannante et al. 1988; Halfar et al. 2006; Westphal et al. 

2010; Michel et al. 2011; Reijmer et al. 2012; Klicpera et al. 

2013; Reymond et al. 2016). The fossil record of encrusting 

cirripedes starts in the Cretaceous (if primitive forms like 

 Archaeochionelasmus Kočí, Newman & Buckeridge, 2017 in 

Kočí et al. 2017 are included), but it is only during the Neo-

gene that barnacles became really frequent in the shallow- 

water environments that they presently master (Darwin 1854; 

Newman et al. 1969; Foster & Buckeridge 1987; Doyle et al. 

1997; Buckeridge 2015).

Worldwide, barnacle facies are particularly well represented 

in the sedimentary sequences of the Neogene and the Quater-

nary (Sakai 1987; Donovan 1988; Kamp et al 1988; Nebelsick 

1989, 1992; Hayton et al. 1995; Doyle et al. 1997; Betzler et 

al. 2000; Nielsen & Funder 2003; Civitelli & Brandano 2005; 

Aguirre et al. 2008; Nomura & Maeda 2008; Radwańska & 

Radwański 2008; Massari & D’Alessandro 2012; Stanton & 

Alderson 2013; Brandano et al. 2015; Buckeridge 2015; 

Buckeridge et al. 2018). Based on the ecology of modern taxa, 

these facies are generally interpreted as shallow-water, 

high-energy, deposits. Although this interpretation is usually 

reasonable, considering the modern distribution of barnacle- 

rich sediments, barnacle facies clearly display a variability 

that reflects environmental differences. The aim of this work is 

to investigate the environmental factors that govern the deve-

lopment of barnacle facies, thus gaining further insights useful 

for palaeoenvironmental reconstructions. To achieve this goal, 

four Burdigalian (early Miocene) barnacle facies, from both 

background image

574

COLETTI, BOSIO, COLLARETA, BUCKERIDGE, CONSANI and EL KATEB

GEOLOGICA CARPATHICA

, 2018, 69, 6, 573–592

the Northern and Southern hemisphere, were analysed and 

compared by means of palaeontology, sedimentology and 

petrography, highlighting their differences and their simila-

rities. The studied barnacle-rich skeletal assemblages are 

located in the well studied successions of the Pietra da Cantoni 

Basin in Italy, of the Sommières Basin in France, and of  

the East Pisco Basin in Peru (Fig. 1; Vannucci et al. 1996; 

Bicchi et al. 2006; Reynaud & James 2012; Coletti et al. 2015; 

Bianucci et al. 2018; DeVries & Jud 2018; Di Celma et al. 

2018b). Previous research provides a firm basis for the inter-

pretation of the barnacle facies of these basins, which has 

received limited attention until now. These facies, despite 

 having in common abundant remains of barnacles, are charac-

terized by different skeletal assemblages and different 

 petrographic composition, thus suggesting different palaeoen-

vironmental settings. The results of this analysis are integrated 

with the existing information on both modern and fossil bar-

nacle facies, to provide a general framework for this kind of 

sedimentary rock.

Geological setting

Pietra da Cantoni Basin, Northern Italy

The Pietra da Cantoni Group was deposited in the eastern 

Monferrato, a part of the Tertiary Piedmont Basin that evolved 

from the late Eocene to the late Miocene, over the inner part of 

the Alpine wedge (Novaretti et al. 1995; Rossi et al. 2009). 

During the Aquitanian, the deformation caused by the rotation 

of the orogenic wedge uplifted the eastern Monferrato and 

resulted in the deposition of the Burdigalian to early Langhian 

limestones of the Pietra da Cantoni (Clari et al. 1995; Novaretti 

et al. 1995; Maffione et al. 2008). The group is divided into 

two depositional sequences (Bicchi et al. 2006). The oldest 

(“Sequence 1” sensu Bicchi et al. 2006) is related to the first 

and localized marine transgression. The youngest (“Sequence 

2” sensu Bicchi et al. 2006) accumulated at the beginning of  

a transgressive trend in the area that lasted for most of  

the Miocene. The second sequence is divided into two units. 

The Lower Unit of Sequence 2 is characterized by coral-

line-algal-nodule rudstones and floatstones interbedded with  

grainstones and rudstones rich either in large benthic fora-

minifera or in barnacles. Deposition occurred during 

 

the Burdigalian (Novaretti et al. 1995; Ruffini 1995; D’Atri et 

al. 1999, 2001). The Upper Unit of Sequence 2 is characte-

rized by foraminiferal oozes, testifying hemipelagic sedimen-

tation. A bed of condensed sediments, rich in glauconite and 

phosphates, separates the two units (Schüttenhelm 1976; 

Bicchi et al. 2006). This interval is related to a period of major 

sediment starvation, caused by the drowning of the carbonate 

factory (Coletti et al. 2015). Both units of Sequence 2 occur in 

the outcrop of Uviglie (Fig. 1; 45°04’42” N, 08°24’48” E), 

where the barnacle-rich facies has been investigated.

Sommières Basin, Southern France

The Alpine Molasse Basin was a seaway during the early 

Miocene; it was about 100 km wide and 1000 km long and 

connected the Western Mediterranean with the Western 

Fig. 1. Location of the studied facies. A — World map including the location of the studied areas highlighted in the panels. B — Western 

Europe, magnification of panel B included in panel A. C — Location of the Sommières Basin (France) and of the Pietra da Cantoni Basin 

(Italy). D — Location of the outcrop of Uviglie (Italy). E — Location of the Boisseron outcrop (France). F — North-western South America, 

magnification of panel F included in panel A. G — Location of the Ullujaya outcrop (Peru).

background image

575

MIOCENE BARNACLE FACIES FROM EUROPE AND SOUTH AMERICA

GEOLOGICA CARPATHICA

, 2018, 69, 6, 573–592

Paratethys (Allen et al. 1985; Rögl 1998; Dercourt et al. 2000; 

Reynaud & James 2012). This narrow sea was dominated  

by strong tidal currents created by the amplification of  

the Atlantic tide entering the basin from the south-west (Allen 

et al. 1985; Harzhauser & Piller 2007). The Sommières Basin 

was a small embayment within the Alpine Molasse Basin, 

located near the junction of the seaway with the Mediterranean 

and connected with the former through a flooded valley 

(Reynaud & James 2012). Within the Sommières Basin, three 

main units are recognized: the Sandy Molasse, the Sandy 

Marls and the Calcareous Molasse (Demarcq 1970; Reynaud 

& James 2012). These units represent two different Burdigalian 

depositional sequences (Berger 1974; Reynaud & James 

2012). The first sequence records the marine transgression in 

the Sommières Basin: in particular, the Sandy Molasse Unit 

represents the transgressive system tract, while the Sandy 

Marls Unit is the highstand system tract (Reynaud & James 

2012). The end of the deposition of the Sandy Marls Unit is 

followed by an abrupt fall in sea level. The subsequent trans-

gression is represented by the Calcareous Molasse Unit, which 

is included in the second sequence (Reynaud & James 2012). 

The Sandy Molasse Unit, which deposited during the early 

Burdigalian, is characterized by limestones composed mainly 

of bryozoans, molluscs, barnacles, echinoids and coralline 

algae (Berger 1974; Reynaud & James 2012). The barna-

cle-rich assemblages occur in the lowermost bed of the unit 

that onlaps the basement of the basin (“Sub-facies A1” sensu 

Reynaud & James 2012). The architecture of these deposits  

is especially clear in the outcrop of Boisseron (Fig. 1;  

43°45’42” N, 04°04’54” E), investigated in the present paper.

East Pisco Basin, South-western Peru

The East Pisco Basin was a Cenozoic semi-enclosed 

forearc-embayment, protected by an archipelago of islands, 

located on the southern coast of Peru (DeVries & Jud 2018;  

Di Celma et al. 2018b). It has been mainly investigated for its 

diverse and exceptionally-preserved Neogene fossil verte-

brates (including pinnipeds, sharks, crocodiles, seabirds, tur-

tles and bony fish) that characterize several outcrops west of 

the Ica River (e.g., Bianucci et al. 2015, 2016 a, b, and previous 

references  therein;  Lambert  et  al.  2014,  2015,  2017 a, b; 

Landini et al. 2017 a, b, 2018; Marx et al. 2017; Gioncada et  

al. 2018). The sedimentary successions were first described  

in the 1990s by Dunbar et al. (1990) and DeVries (1998).  

Within the Palaeogene succession, these authors recognized  

the Caballas (middle Eocene), Paracas (middle to late Eocene 

age) and Otuma (late Eocene to early Oligocene age) forma-

tions. These are followed by the Neogene Chilcatay and Pisco 

formations. The lower Miocene Chilcatay Formation (investi-

gated over the years by Wright et al. 1988; DeVries & Schrader 

1997; León et al. 2008; DeVries & Jud 2018) is 

 

a focus of this paper. Recent surveys in the Western side of  

the Ica River Valley have recognized within the Chilcatay 

Formation distinct depositional sequences, separated by basin-

wide unconformities (Di Celma et al. 2017, 2018a). Based on 

that, the Chilcatay Formation has been divided into the Ct1 

and Ct2 allomembers (Di Celma et al. 2018b). Ct1 is further 

subdivided into two facies associations: Ct1a, composed of 

sandstones and conglomerates (less common than sandstones) 

alternating with siltstones, and Ct1b, including clinobedded 

coarse-grained mixed siliciclastic/bioclastic arenites. Both 

facies associations are characterized by abundant barnacle 

remains. The overlaying Ct2 deposits include massive and 

intensely bioturbated sandstones, changing upwards into mas-

sive siltstones with dolomitized mudstone layers. A tephra 

layer near the top of this sequence locates the Chilcatay 

Formation in the Burdigalian — which agrees with diatom  

and silicoflagellate biostratigraphy (Di Celma et al. 2017, 

2018b). These allomembers are well exposed at Ullujaya  

(Fig. 1; 14°35’06” S, 75°38’30” W), and are investigated in  

the present work.

Material and methods

For the purposes of the present work, skeletal assemblages 

where barnacle remains dominate or codominate the bioclastic 

fraction of the rock are regarded as barnacle facies (with 

codominance meaning the situation in which barnacles, within 

reasonable confidence limits, are equally abundant to another 

group of skeletal grains). Barnacle facies were studied in  

the outcrops of Uviglie, Boisseron and Ullujaya in order to 

describe their macroscopic texture and sedimentary structures 

(Fig. 1). Special attention was given to the observation of fos-

sil barnacle assemblage

s

, focusing on their first-order taxo-

nomic composition. Barnacle preservation and distribution 

were investigated following the methods of Doyle et al. 

(1997), Nomura & Maeda (2008) and Nielsen & Funder 

(2003), which are based on the fragmentation of shells and 

whether or not the specimens are in life position. Representative 

rock samples from each facies were collected for petrographic, 

palaeontological and mineralogical analyses. A large number 

of barnacle specimens were also collected in the studied out-

crops (including isolated opercula, shell fragments, complete 

shells and multi-individual aggregates; Fig. 2). In most cases, 

cirripede palaeontology relies on an analysis of shell frag-

ments as the co-embedding of adjacent valves and, even more, 

the occurrence of complete shells is uncommon (Foster & 

Buckeridge 1987). For this reason, barnacle taxa were studied 

by integrating information from both complete specimens 

(where available) and plates retrieved from the embedding 

rock. From each facies, about 500 g of rock sample was disag-

gregated through freezing-thawing cycles and then wet-sieved 

to isolate barnacle opercula and wall plates (following Aguirre 

et al. 2008). After studying them under a stereomicroscope, 

some selected specimens (67, including compartment frag-

ments, isolated wall plates and complete shells) were prepared 

as thin sections to observe the internal microstructure of  

the shells (Fig. 2), which can be useful for taxonomic identifi-

cation (Cornwall 1956, 1958, 1959, 1960, 1962; Davadie 1963; 

Newman et al. 1969; Newman & Ross 1971; Buckeridge 1983). 

background image

576

COLETTI, BOSIO, COLLARETA, BUCKERIDGE, CONSANI and EL KATEB

GEOLOGICA CARPATHICA

, 2018, 69, 6, 573–592

In particular, attention was paid to interlaminate figures (i.e.

“lames épithéliales” sensu Davadie 1963) that are observable 

in thin sections of the wall plates of the encrusting barnacles 

with a calcareous basis. These interlaminate figures (e.g.,  

Fig. 2) reflect the patterns made by the organic matrix during 

the formation of the calcareous denticles at the basal ends  

of the longitudinal septa of the wall plates that interdigitate 

with the corresponding radial septa of the calcareous basis 

(Newman et al. 1967). As a result of the uncertainties inherent 

in fossil barnacle taxonomy, it was decided to simply differen-

tiate the different taxa occurring in the investigated facies. 

When enough information for a reliable diagnosis was avai-

lable an identification is proposed; otherwise only a provi-

sional identification is attempted. The systematic taxonomy 

used for the classification mostly follows Buckeridge (1983), 

Zullo (1992) and Newman (1996).

The petrographic characteristics of the rocks and their 

 ske letal assemblages were analysed on 35 polished thin 

 sections. The different components were identified and quan-

tified using point-counting method with a minimum of 400 

points per section (Flügel 2010). X-ray powder diffraction 

analysis (XRD) method was used to estimate the carbonate 

and siliciclastic fractions of the whole rock. Samples were first 

ground in an agate mortar, and then mounted on zero-back-

ground silicon plates. The measurements were made with  

a Philips PW1140 diffractometer equipped with CoKα radia-

tion (Kα1 wavelength 1.789 Å) operating at 40 kV and 20 mA. 

Each sample was scanned between 3° and 70° 2θ with a step 

size of 0.02° 2θ and an acquisition time of 1 s per step. X-ray 

pattern treatment was carried out with Panalytical X’pert 

HighScore Plus to identify the mineralogical phases and  

a semiquantitative analysis was carried out using the Reference 

Intensity Ratio (RIR

)

 method (Chung 1974).

Results

Barnalgal, Pietra da Cantoni Group

Skeletal assemblage and mineralogical composition

The barnacle facies of the Pietra da Cantoni Group consists 

of a massive rudstone with a slight yellowish to pinkish colour 

(Fig. 3A–E). This peculiar colour of the rock is partly due to 

the large number of barnacle plates that still retains their pin-

kish pigmentation. The rock is very porous, poorly lithified 

and, except for some rare burrows (Fig. 3C), does not present 

macroscopically evident sedimentary structures; only locally 

flat skeletal elements exhibit a preferential orientation  (Fig. 3F). 

Besides barnacle plates, coralline-algal nodules (rhodoliths) 

are abundant, especially towards the top of the interval. Most 

of them are encrusted by barnacles (Fig. 3D). Sectioned rhodo-

liths show that barnacles are also present within the nodules, 

alternating with layers of coralline algae. Thin section analysis 

shows that the skeletal assemblage is dominated by barnacles 

and coralline algae (mainly Hapalidiales; Table 1; Fig. 3F–G). 

For this assemblage, the new term “barnalgal” is here pro-

posed, complementing the assemblages introduced by Hayton 

et al. (1995). Bryozoans are relatively common (Table 1); their 

association includes branched colonies, globular colonies  

and disarticulated elements of the articulated bryozoan 

Bifissurinella lindenbergi Keij, 1969. Benthic foraminifera are 

present but less abundant (Table 1; Fig. 3F–H), the most com-

mon genera being Amphistegina d’Orbigny, 1826, Elphidium 

Montfort, 1808, and Cibicides Montfort, 1808. Sphaerogypsina 

Galloway, 1933, Miogypsina Sacco, 1893, Operculina  

d’Orbigny, 1826, Nephrolepidina Douvillé, 1911, Eponides 

Montfort, 1808, and Neoconorbina Hofker, 1951 are less com-

mon; Stomatorbina Dorreen, 1948, textulariids and miliolids 

are rare. Echinoids are uncommon in the skeletal assemblage 

and mostly occur as loose spines (Table 1). Molluscs are rare, 

those present being mainly pectinids; serpulids and ostracods 

are very rare (Table 1). The siliciclastic fraction is almost 

non-existent, except for some rare pebbles and some quartz 

grains (Table 1). XRD results support this observation, indi-

cating >> 95 % of carbonate minerals (Table 1).

Barnacle preservation 

Cirripede shells are evenly distributed in the horizon; they 

are generally disarticulated and the plates are slightly abraded 

(Fig. 3E; grade 0 of Nielsen & Funder 2003). Complete shells 

occur on the surface of rhodoliths, together with stubs (i.e.

broken shells of which almost only the base is preserved) and 

partially broken specimens (Figs. 3D; 4A–B). These speci-

mens have randomly-oriented openings, suggesting that they 

are no longer in life position (Type B preservation of Nomura 

Fig. 2. Generalized structure of a balanomorph barnacle shell. Lower 

left corner present a model of a complete shell, partially modified 

after Buckeridge (1983); upper left includes details of the opercular 

plates; lower right includes details of a carinal plate and its internal 

structure revealed by a transverse section above the basal margin 

showing the interlaminate figures, one being enlarged in the upper 

right.

background image

577

MIOCENE BARNACLE FACIES FROM EUROPE AND SOUTH AMERICA

GEOLOGICA CARPATHICA

, 2018, 69, 6, 573–592

& Maeda 2008; displaced clusters of Doyle et al. 1997).  

The specimens are moderately well-preserved, retain their 

colour and only lack the loose and tiny opercula (Fig. 4B).  

The latter are quite abundant in the rock, commonly slightly 

abraded, even though some of them still preserve a pinkish 

colour; sometimes they present gastropod predation holes. 

Scuta and terga do not occur in equal abundance, the scuta 

being much more common than the terga (Fig. 4C–D).

Barnacle identification

The shells are generally small, with the basal diameter in 

adult specimens ranging from slightly under 0.5 cm to a little 

over 1 cm (Fig. 4B). They are comprised of six mural plates 

plus a calcareous, tubiferous basis. Externally the parietes pos-

sess strong longitudinal ribs. Where the colour is preserved, 

the plates are light pink with white spots organized in 

Fig. 3. Barnacle coralline assemblage (barnalgal) of the Lower Unit of Sequence 2 of Pietra da Cantoni Group (Uviglie, Italy). A — Simplified 

stratigraphic column of the barnalgal of the Pietra da Cantoni Group. B — Overview of the Uviglie outcrop including a simple stratigraphic 

column highlighting the different units and the barnalgal facies, BRG = barnalgal assemblage, Sq = sequence. C — Detail of a burrowing trace, 

black arrow = trace. D — Barnacle-encrusted rhodolith, black arrow = barnacles. E — Detail of the texture of the facies, black arrow = Operculina

F — Thin section of a sample of lower part of the barnalgal facies. G — Thin section of a sample of the upper part of the barnalgal facies.  

H — Axial section of a Miogypsina specimen.

background image

578

COLETTI, BOSIO, COLLARETA, BUCKERIDGE, CONSANI and EL KATEB

GEOLOGICA CARPATHICA

, 2018, 69, 6, 573–592

longitudinal rows. Interlaminate figures are present; they have 

short and almost straight main axes with few transverse 

branches (Fig. 4E–F). The terminations of the branches are 

crescent-moon to arrow-shaped. The scuta are 

thick and triangular in shape, with the outer sur-

face characterized by transverse growth lines 

(Fig. 4C). The terga are much thinner, with  

a spur as long as one third of the total length of 

the plate, removed from the basiscutal angle by 

less than half its width, and with a very shallow 

furrow (Fig. 4D). The size and the morphology 

of the recovered opercula indicates that they all 

belong to a single species. This consistency is 

also reflected in the wall plates and their inter-

laminate figures, leading us to conclude that 

only one barnacle species is present in this 

facies. The morphology of the opercula, wall 

plates and interlaminate figures (Fig. 4E–F) 

conforms to a taxon of Balanidae closely allied 

to the amphibalanine Amphibalanus amphitrite 

(Darwin 1854).

Barnamol, Sandy Molasse Unit

Skeletal assemblage and mineralogical 

composition 

This facies consists of rudstone organized in low-angle 

cross-bed sets (Fig. 5A–C). The rock is porous, especially at 

the base of the succession, and exhibits a clear fabric due to 

the common orientation of flat skeletal elements (Fig. 5D–F). 

Barnacle plates and fragments of molluscs are very common. 

Granule-sized bryozoan colonies are also visible on the rock 

surface. According to point-counting analysis the skeletal 

assemblage is dominated by barnacles and molluscs (conse-

quently the assemblage was classified as barnamol sensu 

Hayton et al. 1995; Table 1; Fig. 5 D–I). Barnacles are espe-

cially important at the base of the interval, where they account 

for most of the skeletal grains. Molluscs are more important 

upwards, where they are locally more abundant than barna-

cles; they are mainly represented by ostreids and subordinated 

pectinids. Echinoids are abundant, occurring as spines and test 

fragments (Table 1). Bryozoans are common (both branching 

and globular colonies are present; Table 1). Coralline algae 

and benthic foraminifera are very rare, the latter being mostly 

represented by Cibicides,  Amphistegina and Miogypsina 

(Table 1). The siliciclastic fraction is important and accounts 

for 16.5 % of the detected elements (Table 1; Fig 5D–L).  

It includes quartz grains and fragments of sedimentary rocks 

(mainly limestones and sandstones). XRD results highlight  

the pre sence of about 30 % of silicate minerals and 70 % of 

carbo nates (Table 1).

Barnacle preservation

Wall plates are invariantly disarticulated and evenly distri-

buted within the rock (Type D of Nomura & Maeda 2008; 

comminuted shell bed of Doyle et al. 1997). They are frag-

mented, disarticulated and heavily abraded, often exposing 

Pietra da Cantoni 

Group

Sandy Molasse 

Unit

Chilcatay Formation

Barnalgal

Barnamol

Ct1a  

barnacle facies

Ct1b  

barnacle facies

Petrographic composition [Point counting]
Bioclastic components

95.0 %

68.5 %

22.0 %

24.0 %

Terrigenous components

0.5 %

16.5 %

65.0 %

30.5 %

Sparite

4.0 %

8.0 %

10.0 %

41.0 %

Micrite

0.5 %

7.0 %

3.0 %

4.5 %

Detail of the bioclastic fraction [Point counting]
Barnacles

41.5 %

34.5 %

71.5 %

80.0 %

Molluscs

0.5 %

40.5 %

8.5 %

7.5 %

Echinoids

1.0 %

16.5 %

18.5 %

8.0 %

Bryozoans

6.5 %

8.0 %

< 0.5 %

///

Coralline algae

47.0 %

< 0.5 %

///

///

Benthic foraminifera

3.5 %

< 0.5 %

1.0 %

4.0 %

Serpulids

< 0.5 %

< 0.5 %

< 0.5 %

< 0.5 %

Ostracods

< 0.5 %

///

///

///

Mineralogical composition [XRD]
Carbonates

97.0 %

72.0 %

16.0 %

31.0 %

Silicates

3.0 %

28.0 %

71 %

64.0 %

Salts

0.0 %

0.0 %

13.0 %

5 %

Table 1: Petrographic composition, skeletal assemblage and mineralogical content of 

the examined barnacle facies.

Fig. 4. Barnacles of the barnalgal of Uviglie (Italy). A — A rhodolith 

completely encrusted by barnacles (cf. Amphibalanus sp.), upper part 

of the Lower Unit of Pietra da Cantoni Sequence 2 (Colma, Fig. 1).  

B — Barnacle shells (cf. Amphibalanus sp.) growing on a rhodolith.  

C — Internal and external surfaces of the same scutal plate.  

D — Internal and external surfaces of the same tergal plate.  

E — Interlaminate figures of a wall plate of cf. Amphibalanus sp.  

F — Detail of the interlaminate figures of the plate depicted in panel E.

background image

579

MIOCENE BARNACLE FACIES FROM EUROPE AND SOUTH AMERICA

GEOLOGICA CARPATHICA

, 2018, 69, 6, 573–592

their internal structure (Grade 1 and 2 of Nielsen & Funder 

2003). None of the observed plate fragments preserve their 

original colour. Very rare scuta, but no terga, were recovered 

from the disaggregated rock sample. They are fragmented, 

abraded and exhibit pitted surfaces probably due to incipient 

pressure solution.

Barnacle identification

The preservation state of the cirripede remains greatly 

 hinders identification based on their macroscopic features. 

The inner structure of the plates and the interlaminate figures 

indicates that three different taxa are probably present: 

Concavinae? indet., Balaninae gen. et sp. indet. 1, and 

Balaninae gen. et sp. indet. 2 (Fig. 6). The presence of more 

than one barnacle taxon is also supported by the scutal plates, 

which exhibit at least two different morphologies: one charac-

terized by transverse growth lines only and another exhibiting 

both transverse growth lines and longitudinal (i.e., radial) 

striae (Fig. 6F–G, respectively). The specimens identified as 

Concavinae? indet. have small parietal-tubes, sometimes 

divided by septa and characterized by interlaminate figures 

with a straight axis and few transverse branches with crescent- 

moon- to arrow-shaped terminations (Fig. 6A). The presence 

Fig. 5. Barnacle-mollusc assemblage (barnamol) of the Sandy Molasse Unit of the Sommières Basin (Boisseron, France). A — Simplified 

stratigraphic column of the barnamol facies of the Sandy Molasse Unit. B — Overview of the Boisseron outcrop. C — Macroscopic texture of 

the facies. D — Thin section of a sample from the barnamol facies with common orientation of the bioclasts. E — Thin section of a barna-

cle-rich sample of the barnamol facies. F — Thin section of a mollusc-rich sample of the barnamol facies. G — Detail of a Miogypsina speci-

men. H — Detail of a cross-section of a tubiferous barnacle wall plate exhibiting interlaminate figures.

H

background image

580

COLETTI, BOSIO, COLLARETA, BUCKERIDGE, CONSANI and EL KATEB

GEOLOGICA CARPATHICA

, 2018, 69, 6, 573–592

of representatives of the subfamily Concavinae is as also sug-

gested by the occurrence of scuta with both longitudinal striae 

and transverse growth lines; however, it must be stressed that 

the latter feature is also observed in the extant amphibalanine 

Amphibalanus eburneus (Gould, 1841). Concavine barnacles 

are known to occur in the Neogene deposits of the Alpine 

Molasse Basin. De Alessandri (1906) pointed out the presence 

of “Balanus concavus Bronn, 1831 (=Concavus concavus) in 

lower Miocene strata of the Upper Marine Molasse of 

Switzerland, but under this specific name several taxa of spe-

cific and generic rank are lumped (Newman 1982; Zullo 1992; 

Carriol 2000). More recently, concavine barnacles (belonging 

to the genus Chesaconcavus Zullo, 1992) have been reported 

by Carriol & Schneider (2008) and Carriol & Menkveld-

Gfeller (2010) from Burdigalian beds of the Upper Marine 

Molasse of Bavaria and Switzerland, respectively. The speci-

mens referred to Balaninae gen. et sp. indet. 1 have larger 

 parietal tubes and interlaminate figures with a long axis and 

numerous transverse branches with club-shaped terminations 

(Fig. 6B–C). Similar interlaminate figures have been figured 

by Nebelsick (1989: pl. 4, figs. 2–3) for Burdigalian barnacles 

of the Zogelsdorf Formation of Austria (Alpine Molasse 

Basin) and by Davadie (1963; pl. XIV) for recent specimens 

of the subfamily Balaninae. The specimens belonging to 

Balaninae gen. et sp. indet. 2 have large tubes and interlami-

nate figures with a bent axis and numerous transverse branches 

with club-shaped terminations; differing from the other 

groups, these branches present complex ramifications 

 

(Fig. 6D–E).

Barnacle facies, Chilcatay Formation

Skeletal assemblage and mineralogical composition

Two slightly different barnacle facies can be recognized in 

Ct1a and Ct1b facies associations, respectively (Fig. 7A–B).

The barnacle facies of Ct1a consists of poorly-sorted 

coarse-sandstones to conglomerates with siliciclastic and bio-

clastic elements (Fig. 7A, C–E); the barnacle-bearing horizons 

alternate with medium- to fine-grained sandstones, almost 

devoid of bioclasts (Fig. 7A).The rock is massive, porous and 

poorly cemented. Barnacles are very common, occurring as 

clusters of shells, isolated individuals and large shell frag-

ments (Fig. 7D–E).According to thin section analysis, barna-

cles dominate the skeletal assemblage (Table 1; Fig. 7F–G). 

Echinoids are also very common; molluscs are less abundant 

and mainly represented by ostreids and pectinids (Table 1). 

Benthic foraminifera are very rare (mainly Cibicides

Peneroplis  Montfort, 1808 and Nonion  Montfort, 1808; 

 

Table 1). Serpulids are uncommon in the skeletal assemblage, 

but occasionally they occur in large clusters. Bryozoans are 

very rare (Table 1). The siliciclastic fraction is important and 

accounts for more than half of the components (Table 1;  

Fig. 7F–G). It consists of fragments of igneous rocks from  

the basement and ash-flow tuffs. XRD results indicate > 70 % 

of silicate minerals and only lesser amounts of carbonates and 

salts (mainly gypsum; Table 1).

The barnacle facies of Ct1b is organized into clinobeds of 

well-sorted, coarse-grained mixed siliciclastic-bioclastic 

deposits, which are more cemented and less porous than those 

of the lower facies (Fig. 7A). On a macro-scale at the outcrop, 

barnacles dominate the bioclastic fraction (occurring as clus-

ters, isolated individuals and shell fragments), but bivalves 

also occur (Fig. 7 H–J). In thin section, barnacle remains are 

the most abundant group of skeletal grains; however, the com-

mon presence of sparite-filled molds suggests that molluscs 

were also important (Table 1; Fig. 7 K–L). Among the few 

preserved bivalve specimens, ostreids and pectinids are domi-

nant. Echinoids are also abundant (Table 1). Benthic fora-

minifera are less common (mainly Cibicides and Peneroplis

Table 1). Rare calcareous tubes also occur (Table 1). Based on 

point-counting analyses, siliciclastic components are slightly 

more important than bioclasts and they mainly consist of 

 pebbles and granules of igneous basement rocks (Table 1;  

Fig. 7F–G, K–L). XRD results suggest that silicate minerals 

account for > 60% of the rock, whereas carbonates only repre-

sent 30 % (Table 1). Minor amounts of gypsum are also  present 

(Table. 1).

Fig. 6. Barnacles of the barnamol of Boisseron (France). 

 

A —  Concavinae? indet., interlaminate figures; note the straight axis, 

few branches and crescent-moon to arrow-shaped terminations.  

B —  Balaninae gen. et sp. indet. 1, interlaminate figures; note the 

long axis and numerous branches with club-shaped terminations.  

C —  Balaninae gen. et sp. indet. 1, detail of the interlaminate figures. 

D —  Balaninae gen. et sp. indet. 2, interlaminate figures; note  

the bent axis and numerous, ramified branches with club-shaped 

 terminations.  E —  Balaninae gen. et sp. indet. 2, detail of the inter-

laminate figures. F —  External surface of a scutal plate exhibiting 

transverse growth lines only. G —  External surface of a scutal plate 

exhibiting both longitudinal striae and transverse growth lines.

background image

581

MIOCENE BARNACLE FACIES FROM EUROPE AND SOUTH AMERICA

GEOLOGICA CARPATHICA

, 2018, 69, 6, 573–592

Fig. 7. Barnacle facies of the Chilcatay Formation (Ullujaya, Peru). A —  Simplified stratigraphic column of the barnacle facies of  

the Chilcatay Formation. B —  Overview of the Ullujaya outcrop including a simple stratigraphic column highlighting the different facies.  

C —  Macroscopic texture of the barnacle facies of Ct1a. D —  Detail of the texture and clast composition of the Ct1a barnacle facies.  

E —  Detail of the texture and composition of the Ct1a barnacle facies. F —  Thin section of a sample of the Ct1a barnacle facies. G —  Detail 

of sample of the Ct1a barnacle facies with barnacles and foraminifera. H —  Macroscopic texture of the Ct1b barnacle facies. I —  Detail of 

the texture and bioclastic composition of the Ct1b barnacle facies. J —  Detail of the texture and composition of the Ct1b barnacle facies.  

K —  Thin section of a sample of the Ct1b barnacle facies. L —  Detail of the abundant sparite-filled molds in the Ct1b barnacle facies.

background image

582

COLETTI, BOSIO, COLLARETA, BUCKERIDGE, CONSANI and EL KATEB

GEOLOGICA CARPATHICA

, 2018, 69, 6, 573–592

Barnacle preservation

In both facies, the cirripede remains occur as displaced 

 clusters, displaced complete shells and wall plate fragments 

(Fig. 8). Unlike the barnacle-coralline facies, the clusters are 

generally detached from their original substrate (Type C of 

Nomura & Maeda 2008; displaced clusters of Doyle et al. 

1997).

Complete specimens are generally moderately well-pre-

served, but display evidence of abrasion and lack the opercula 

(Fig. 8). The best preserved specimens retain their pigmen-

tation (Fig. 8A, D). The shells are often filled by sand.  

In the barnacle facies of Ct1a, this filling is remarkably rich in 

bioclastic fragments and has less abundant mineral grains than 

the rock embedding the specimens. Disarticulated plates are 

often abraded (Grade 1 of Nielsen & Funder 2003). Opercula 

are rare — only a couple of abraded and fragmented scuta 

have been recovered.

Barnacle identification

Based on macroscopic features of the wall plates and on 

their internal microscopic structure, three barnacle taxa have 

been recognized: cf. Austromegabalanus sp., Balanidae indet., 

and Concavinae indet. (Fig. 8). The specimens identified as cf. 

Austromegabalanus  sp. have a shell diameter approaching  

4 cm (Fig. 8A–C); they are comprised of six wall plates with 

broad, well-developed radii plus a calcareous, tubiferous, 

basis (Fig. 8A). The best preserved specimens are longitudi-

nally striped, with pink-purple bands alternating with white 

(Fig. 8A). In thin section, the radii are tubiferous and their 

sutural edges bear transversely oriented septa with secondary 

Fig. 8. Barnacles of the barnacle facies of Ullujaya (Peru). A — cf. Austromegabalanus sp., complete shell (from Ct1a). B — cf. Austro­

megabalanus  sp., interlaminate figures (from Ct1a). C — cf.  Austromegabalanus  sp., detail of the interlaminate figures (from Ct1a).  

D — Balanidae indet., cluster of shells (from Ct1a). E — Balanidae indet., interlaminate figures (from Ct1a). F — Balanidae indet., detail of 

the interlaminate figures (from Ct1a). G — Concavinae? indet.,cluster of shells (from Ct1b). H — Concavinae? indet., interlaminate figures 

(from Ct1a). I — Concavinae? indet., detail of the interlaminate figures (from Ct1a).

background image

583

MIOCENE BARNACLE FACIES FROM EUROPE AND SOUTH AMERICA

GEOLOGICA CARPATHICA

, 2018, 69, 6, 573–592

denticles on the lower side only (a diagnostic character of  

the megabalanine genera included in the tribe Austro-

megabalanini; Newman 1979; Buckeridge 2015). The interla-

minate figures present an almost straight main axis with many 

transverse branches characterized by multiple arrow-shaped 

terminations (Fig. 8B–C). Austromegabalanus Newman, 1979 

is known to occur along the coasts of southern Peru and in  

the Miocene sediments of the Pisco Basin (e.g., Newman 

1991; Carriol et al. 1987). The specimens referred to Balanidae 

indet. are medium- to large-sized shells, comprised of six wall 

plates and characterized by a longitudinally striated outer wall 

with spiny-bulging protuberances creating a rough and irre-

gular external surface (Fig. 8D–F). Some individuals display  

a pink colour on the plates (Fig. 8D). In the lowermost portion 

of the shell the basal edges of the wall plates become inflected, 

and grow horizontally inwards, thus producing an inward- 

tapering calcareous membrane. This feature is often incom-

plete (i.e., a hole is present at the centre of the basal calcareous 

membrane). In juvenile individuals that were overgrown by 

adults the basal calcareous membrane is not developed.  

The parietes of the wall plates are porous and multiple rows  

of tubes are present (Fig. 8E). Large pores occur also in  

the sheath (Fig. 8E). The radii are solid. Interlaminate figures 

are arborescent, with a long straight axis and short transverse 

branches with crescent-moon to kidney-shaped terminations 

(Fig. 8F). Based on the general architecture of the six-plated 

shell and the presence of complex interlaminate figures, these 

specimens are provisionally identified as indeterminate 

 balanids. Among Balanidae, a porous sheath characterizes 

Titobustillobalanus tubutubulus Carriol & Álvarez-Fernández, 

2015 from the latest Pleistocene of Spain and various species 

of the amphibalanine genus Fistulobalanus Zullo, 1984.  

The specimens identified as Concavinae indet. are medium- to 

small-sized, six-plated shells with a basal diameter ranging 

between 1 and 2 cm (Fig. 8G–I). They appear to have been 

strictly cluster-forming organisms, since even displaced single 

individuals show an elongated, tubiferous, basal plate tapering 

downwards (a morphology that is typically associated with  

a gregarious growth habit, e.g., Newman & Ross 1976).  

Some specimens display a faint pink to purple pigmentation. 

The wall plates are porous and display a single row of tubes 

divided by frequent transverse septa (Fig. 8H). Interlaminate 

figures have a long and straight axis with numerous, closely- 

spaced transverse branches characterized by arrow-shaped 

terminations (Fig. 8I). These branches are often organized in 

pairs, stemming from the same point of the main axis (Fig. 8I). 

The very rare opercula retrieved in the sieved material may be 

attributed to this group of specimens due to their size. They 

consist of triangular scuta characterized by transverse growth 

lines.

A further, currently unidentified, barnacle taxon was 

observed only once in the Ct1a facies association. It is repre-

sented by small-sized individuals (around 0.5 cm in basal 

diameter) forming a cluster which partially encrusts a pectinid 

shell. These small-sized barnacles could also represent juve-

niles of one of the afore-mentioned taxa. The common 

occurrence of pectinids with small barnacle attachment scars 

(Anellusichnus circularis Santos, Mayoral & Muñiz, 2005) on 

the outer surface could indicate that this type was less uncom-

mon than suggested by body fossils alone.

Discussion

Facies interpretation

Previous works based on coralline-algal and foraminiferal 

assemblages suggest that both the Pietra da Cantoni and  

the Sommières Basin barnacle facies developed in a tropical 

setting (Vannucci et al. 1996; Coletti et al. 2015, 2017, 2018). 

Vertebrate and mollusc assemblages indicate warm-temperate 

conditions for the East Pisco Basin during the deposition of 

the Ct1 beds of the Chilcatay formation (DeVries & Frassinetti 

2003; Bianucci et al. 2018). The observation of the forami-

niferal genus Peneroplis, which is typical of warm and 

warm-temperate water (Murray 2006), supports this hypo-

thesis. The skeletal assemblages observed in all the sites, 

 characterized by the abundance of filter feeding organisms 

and the scarcity of symbiont-bearing taxa, in warm water 

points to a nutrient-rich setting (Brasier 1995 a, b). This is also 

in agreement with other works on the studied successions 

(Dunbar et al. 1990; Reynaud & James 2012; Coletti et al. 

2015, 2017; Bianucci et al. 2018).

Both barnacle facies of the Chilcatay Formation have a “low 

diversity” skeletal assemblage, with barnacle and molluscs 

(mainly epifaunal bivalves like ostreids and pectinids) 

accounting  for  > 75  %  of  the  skeletal  fragments  (Table  1).  

It should be noticed that mollusc abundance was probably 

reduced by selective dissolution during diagenesis; this parti-

cularly applies to the barnacle facies of Ct1b, where mol-

lusc-shaped, sparite-filled molds are common. The abundance 

of barnacles and epifaunal bivalves points toward an exposed 

setting since both groups are favoured by high water energy 

(Farrow et al. 1978; Nebelsick 1992). A setting favourable for 

barnacles is also supported by the diverse barnacle assemblage 

with at least three common taxa. The preservation of barnacles 

in both facies is similar. Whole individuals are present (either 

as displaced clusters or single shells) and always separated 

from their substrate. Some specimens still preserve their pig-

mentation but the majority of them do not. Opercula are 

extremely rare. These characteristics suggest significant 

reworking, compatible with a setting above fair-weather 

wave-base. This hypothesis is in agreement with Bianucci et 

al. (2018) and Di Celma et al. (2018b), which interpreted Ct1b 

as a shoreface deposit, probably formed at a water depth of less 

than 15–20 m (assuming a storm-wave base around 15–20 m 

in accordance with the models of Hernández-Molina et al. 

2000; Massari & Chiocci 2006). The same authors interpreted 

Ct1a as an offshore deposit resulting from the downslope 

transport of shoreface material (below 30–40 m of water depth 

since the height of the clinoforms is about 15–20 m). Since 

both facies have the same signature, it is likely that most of 

background image

584

COLETTI, BOSIO, COLLARETA, BUCKERIDGE, CONSANI and EL KATEB

GEOLOGICA CARPATHICA

, 2018, 69, 6, 573–592

reworking occurred in the shoreface environment, with 

 

the down slope movement only mildly affecting the associa-

tion. This is also supported by the sediment preserved within 

the shells of the Ct1a barnacles, which is similar in composi-

tion to that of Ct1b (more bioclasts and less siliciclastic 

particles).

The skeletal assemblage of the barnamol facies of the Sandy 

Molasse Unit is also overwhelmingly dominated by barnacles 

and epifaunal bivalves, suggesting a high-energy setting 

(Table 1; Farrow et al. 1978; Nebelsick 1992). Extensive evi-

dence of dissolution was not observed in this facies, sugges-

ting that the mollusc abundance was not significantly reduced. 

Suitable conditions for barnacles are suggested by the pre-

sence of a diverse cirripede assemblage. There are no complete 

shells and well-preserved specimens in general are missing; 

opercula are also extremely rare. The poor state of preserva-

tion of the barnacle remains and the cross-bedded structure of 

this facies are consistent with a very shallow and proximal 

marine environment (nearshore zone, less than 15 m of water 

depth). This interpretation is in agreement with previous sedi-

mentological studies on the Boisseron outcrop of the Sandy 

Molasse Unit (Reynaud & James 2012). The environment was 

probably characterized by higher hydrodynamic energy than 

that of the Chilcatay Formation.

The barnalgal of the Pietra da Cantoni significantly diverge 

from the other barnacle facies. Siliciclastic elements are much 

rarer (Table 1), ruling out the presence of nearby rocky cliffs. 

Its skeletal assemblage is more diverse, coralline algae are 

codominant, benthic foraminifera are more diverse, and there 

are fewer molluscs and echinoids than in the other facies 

(Table 1). There is no extensive evidence of dissolution sug-

gesting that the low abundance of molluscs is not a result of 

diagenetic processes (although leaching of aragonite might 

have occurred since gastropods are absent even thought 

 gastropods predation holes are present on barnacles). Unlike 

the other facies only a single taxon of barnacle was recog-

nized. Differing from the other facies, most of the shells are 

well preserved, retain their original colour and are associated 

with their substrate (i.e., the rhodoliths). Opercula are abun-

dant and only limited reworking can be inferred by the unequal 

ratio of terga and scuta (the heavier scuta being more common 

than the lighter terga). These features suggest a less exposed 

environment, below the fair-weather wave base (probably 

between 20 and 40 m of water depth). This is in agreement 

with previous studies that interpreted this material as a short- 

distance mass-transport of inner-middle ramp material depo-

sited in a slightly deeper middle-ramp setting (below 50–60 m 

of water depth; Schüttenhelm 1976; Coletti et al. 2015). Short-

distance transport and rapid burial, without further reworking, 

are strongly supported by the preservation of colour in barna-

cles (Hollingworth & Barker 1991; Aguirre et al. 2008).

The detailed comparison of skeletal assemblages, petro-

graphic characteristics and barnacle preservation and diversity 

clearly separates the barnalgal facies from the remaining three 

facies. The latter have a diverse barnacle assemblage mainly 

associated with epifaunal bivalves and echinoids, and 

comprise an important siliciclastic fraction. The barnalgal do 

not. The French and Peruvian facies are related to a very shal-

low, nearshore carbonate factory, developed along a high- 

energy rocky coast. This is the ideal setting for encrusting 

barnacles; they are perfectly adapted to systems where hard 

surfaces are abundant. The environment of the barnalgal facies 

deviates substantially from this optimum. There are less avai-

lable surfaces and most of them are coralline algae, which 

directly compete with barnacles for space (as suggested by  

the presence of alternating layers of barnacles and coralline 

algae in the rhodoliths). This might explain why various dif-

ferent species of barnacles are found in the first group of facies 

and only a single species characterizes the barnalgal facies.

Environmental controls on barnacle facies

Modern barnacle facies occur from the Poles to the Equator 

and, notwithstanding this large latitudinal variation, they show 

clear similarities (see Table 2 for the complete references list 

and Fig. 9 for the locations). Barnacle facies usually contain 

abundant molluscs (Table 2; Hoskin & Nelson 1969; Milliman 

1972; Müller & Milliman 1973; Farrow et al. 1978; Scoffin 

1988; Wilson 1988; Halfar et al. 2006; Westphal et al. 2010; 

Michel et al. 2011; Reijmer et al. 2012; Reymond et al. 2016). 

Echinoids are important contributors in most of the occur-

rences (Table 2; Hoskin & Nelson 1969; Farrow et al. 1978; 

Scoffin 1988; Wilson 1988; Michel et al. 2011). At low lati-

tudes, the association with hermatypic corals is also possible 

(Table 2; Glynn & Wellington 1983; Halfar et al. 2006; 

Reymond et al. 2016). Coralline algae and benthic foramini-

fera are practically absent (Table 2; Hoskin & Nelson 1969; 

Farrow et al. 1978; Scoffin 1988; Wilson 1988; Halfar et al. 

2006; Westphal et al. 2010; Michel et al. 2011; Reijmer et al. 

2012; Reymond et al. 2016). Barnacle facies have an impor-

tant siliciclastic fraction, which accounts for at least 10 % of 

the grains, although it is typically much higher (Table 2; 

Hoskin & Nelson 1969; Milliman 1972; Müller & Milliman 

1973; Farrow et al. 1978; Scoffin 1988; Scoffin and Bowes 

1988; Wilson 1988; Halfar et al. 2006; Westphal et al. 2010; 

Michel et al. 2011; Reijmer et al. 2012; Frank et al. 2014; 

Reymond et al. 2016). As far as the depositional environment 

is concerned, most modern barnacle facies are related to shal-

low-water (less than 50 m and generally less than 20 m), high- 

energy environments with nearby rocky outcrops (Table 2; 

Hoskin & Nelson 1969; Farrow et al. 1978; Scoffin 1988; 

Wilson 1988; Henrich et al. 1995; Halfar et al. 2006; Westphal 

et al. 2010; Michel et al. 2011; Reijmer et al. 2012; Reymond 

et al. 2016). Most barnacle facies are also related to plank-

ton-rich water (Table 2; Müller& Milliman 1973; Taviani et al. 

1993; Henrich et al. 1995; Westphal et al. 2010; Michael et al. 

2011; Reijmer et al. 2012; Klicpera et al. 2013; Frank et al. 

2014; Reymond et al. 2016).

Those occurrences that significantly deviate from this gene-

ral model are located at polar latitudes and in bathyal settings 

(Table 2). The barnacle facies of the Barents Sea are characte-

rized by almost pure carbonates, composed of barnacles and 

background image

585

MIOCENE BARNACLE FACIES FROM EUROPE AND SOUTH AMERICA

GEOLOGICA CARPATHICA

, 2018, 69, 6, 573–592

benthic foraminifera (Henrich et al. 1995). Unlike the exam-

ples discussed in detail herein, carbonate production is loca-

lized far away from the coast, on shallow submerged rocky 

shoals colonized by kelp forests (Henrich et al. 1995). The Ross 

Sea barnacle facies develop on hard substrates under high 

energy conditions; however, they are dominated by Bathy­

lasma corolliforme (Hoek, 1883) that lives in deeper water 

than most other encrusting barnacles (Taviani et al. 1993; 

Frank et al. 2014). Bathylasmatidae-dominated facies occur 

also in even deeper, bathyal, settings (Table 2; Newman & 

Ross 1976; Buckeridge 1999). They are found close to sea-

mounts and, unlike their shallow-water counterparts, they are 

associated with fine grained sediments and deep-sea fauna 

(deep-water corals, bryozoans and planktonic foraminifera; 

Location

Age

Key 

references

Barnacle taxa

Associated bioclasts

Texture

Clastic 

particles

Environmental information

Antartica, Ross 

Sea Shelf

Pleistocene–

Recent

Frank et al. 

2014; Taviani 

et al. 1993

Bathylasma 

corolliforme 

(Hoek, 1883)

Locally only composed by 

barnacles; minor local contributions 

from stylasterine hydrocorals, 

bryozoans and benthic foraminifera; 

rare bivalves and serpulids.

According to Taviani et al. (1993) 

barnacles are mainly associated with 

foraminifera

Coarse to 

very coarse 0 to 40 %

Polar; more than 350 m of 

water depth; high energy; 

plankton-rich water

South Pacific 

Ocean, South 

Tasman Rise

Holocene-

Recent

Buckeridge 

1999

Tetrachaelasma 

tasmanicum 

Buckeridge, 

1999

Mainly solitary deep-water 

scleractinian corals, but also 

planktonic foraminifera 

Very fine to 

coarse

Present, both 

rocks and 

mud

Bathyal environment, 2100-

3600 m of water depth

South 

Madagascar; 

Indian Ocean

Recent

Newman & 

Ross 1976

Tetrachaelasma 

Newman & Ross, 

1971

///

Coarse

///

Bathyal environment, 2000 m 

of water depth

Barents Sea, 

Spitsbergenbank Recent

Henrich et al. 

1995

Balanus 

crenatus 

Bruguière, 1789

Mainly benthic foraminifera

Coarse

Pure 

carbonate

Polar; between 6 and 20 m of 

water depth; high energy; 

plankton-rich water

Alaska, 

Alexander 

Archipelago

Recent

Hoskin & 

Nelson 1969

///

Major contributions from molluscs 

and echinoids; minor contributions 

from benthic foraminifera; rare 

ahermatypic corals, coralline algae 

and bryozoans

Fine to 

coarse

15 % to 60 % Cold sub-polar; macrotidal 

setting

Canada, 

Newfoundland

Recent/Sub-

recent

Müller and 

Milliman 

1973

///

Major contribution from molluscs

Coarse

50 % to 80 %

Cold sub-polar; 60 to 80 m of 

water depth; plankton-rich 

water

USA, locally 

along Carolina 

and Florida 

Atlantic shelves

Recent/Sub-

recent

MacIntyre & 

Milliman 

1970; 

Milliman 

1972

Balanus calidus 

Pilsbry, 1916 

(temptative)

Mainly coralline algae (locally 

codominant) 

///

3 % to 74 %, 

39 % on 

average

Barnacles growing on a relict 

algal ridge that borders the 

upper slope and the outer shelf 

edge (50-150m); strong 

currents.

USA, locally 

along Carolina 

and Florida 

Atlantic shelves

Recent/Sub-

recent

Milliman 

1972

Balanus calidus 

(temptative)

Mainly molluscs  (locally 

codominant) 

///

3 % to 94 %, 

45 % on 

average

Barnacles growing on 

submerged rocky outcrops

U.K., western 

Scotland

Recent

Farrow et al. 

1978; Scoffin 

1988; Wilson 

1988

Semibalanus 

balanoides 

(Linnaeus, 1767)

Codominance with molluscs; 

important contributions from 

echinoids; minor contributions from 

serpulids and coralline algae; rare 

bryozoans and benthic foraminifera. 

According to Wilson (1988) the 

contributions from serpulids can be 

very important

Coarse

0 % to 50 %, 

with most of 

the values 

around 20 %

Cool-temperate; less than 50 m 

of water depth; hard substrate; 

high energy

Mexico, Bahía 

de Los Angeles

Recent

Halfar et al. 

2006

///

Major contributions from molluscs 

and hermatypic corals; minor 

contributions from bryozoans and 

echinoids; rare serpulids; very rare 

coralline algae and barnacles

Fine to 

coarse

30 % to 70 %

Warm-temperate; upwelling 

currents;  high nutrient 

concentration (eutrophic); less 

than 20 m of water depth; hard 

substrate; high energy

Pacific side of 

Panama, Gulf of 

Panama

Recent

Reijmer et al. 

2012

///

Mainly bivalves; common 

gastropods; minor contributions 

from echinoids, serpulids, benthic 

foraminifera and coralline algae

Coarse

At least 10 %

Tropical; upwelling currents; 

high nutrient concentration 

(eutrophic); less than 50m of 

water depth

Mauritania, 

Banc D'Arguin

Recent

Westphal et 

al. 2010; 

Michel et al 

2011; 

Klicpera et al. 

2013

Balanus Costa, 

1778

Codominance with bivalves; 

significant contributions from 

echinoids; rare gastropods and 

benthic foraminifera; very rare 

bryozoans

Medium

30 % to 50 %

Tropical; upwelling currents; 

high nutrient concentration 

(eutrophic); sub-tidal, less than 

20 m of water depth; high 

energy

Chile, Galápagos 

Archipelago

Recent

Glynn & 

Wellington 

1983; 

Westphal et 

al. 2010; 

Reymond et 

al. 2016

Megabalanus 

tintinnabulum 

(Linnaeus, 1758)

[“barnamol”] Important gastropods 

and serpulids; minor contributions 

from bryozoans, bivalves, echinoids 

and coralline algae

Fine to 

coarse

10 % to 70 %

Equatorial; strong upwelling; 

less than 15 m of water depth; 

mesotrophic

[“barnamolcor”] Codominance with 

gastropods; important contributions 

from coralline algae and corals; 

minor contributions from echinoids, 

serpulids, bryozoans and bivalves

10 % to 50 %

Equatorial; moderate 

upwelling; less than 15 m of 

water depth; oligotrophic

Table 2: Sub-recent and recent barnacle facies, including information on the position, barnacle assemblage, skeletal assemblage, texture, 

clastic fraction and available environmental information.

background image

586

COLETTI, BOSIO, COLLARETA, BUCKERIDGE, CONSANI and EL KATEB

GEOLOGICA CARPATHICA

, 2018, 69, 6, 573–592

Newman & Ross 1976; Buckeridge 1999). The barnalgal asso-

ciation described by Milliman (1972) along the South-eastern 

Coast of the United States is also peculiar and differs from 

most of the other barnacle facies. It is related to the recent 

barnacle colonization of a relict algal ridge growing during  

the last glacial period that now borders the outer shelf edge.

Neogene and Quaternary barnacle facies follow the same 

pattern as their modern counterparts (Table 3; Fig. 9). The ske-

letal assemblage is mainly characterized by barnacles and 

molluscs, with common bryozoans and echinoids, rare benthic 

foraminifera and coralline algae; the siliciclastic fraction is 

also important (Table 3; Sakai 1987; Kamp et al. 1988; 

Nebelsick 1989, 1992; Doyle et al. 1997; Betzler et al. 2000; 

Nielsen & Funder 2003; Aguirre et al. 2008; Nomura & Maeda 

2008; Stanton & Alderson 2013; Brandano et al. 2015; 

Buckeridge et al. 2018). They are generally interpreted as 

high-energy shallow-water deposits related to non-oligotro-

phic conditions (Table 3; Buckeridge 1983; Sakai 1987; Kamp 

et al. 1988; Nebelsick 1989, 1992; Doyle et al. 1997; Nielsen 

& Funder 2003; Civitelli & Brandano 2005; Aguirre et al. 

2008; Nomura & Maeda 2008; Stanton & Alderson 2013; 

Brandano et al. 2015).

When considering the distribution of both modern and  

fossil barnacle concentrations, the major factors controlling 

the deve lopment of a barnacle-dominated carbonate factory 

seem to be the presence of suitable hard substrates, hydro-

dynamic energy and nutrient availability. Although barnacle 

accumulations are more common in the cool-temperate and 

cold realms, it seems that the lack of nutrients is the main fac-

tor that limits their distribution in the tropical zone. Therefore, 

an inshore rocky substrate characterized by high-energy con-

ditions and abundant nutrient-supply is very favourable to  

the development of a barnacle-dominated carbonate factory. 

This setting represents an environmental optimum for barna-

cles (Sanford & Menge 2001) and other hard-substrate orga nisms 

(e.g., ostreids and pectinids), and can lead to the formation of 

classical barnamol facies carbonates (sensu Hayton et al. 

1995). Coralline algae might occur in this environment, but 

mostly as thin adherent crusts, with a poor preservation poten-

tial. Foraminifera can also occur, but they prone to be swept 

away by currents and deposited elsewhere (Farrow et al. 1978; 

Nebelsick 1992). The barnacle-rich facies investigated in this 

study comply with this model, except for the barnalgal facies. 

The latter represents a more atypical setting, located in slightly 

deeper water and where the surface available for barnacle 

colo nization is more limited. Facies similar to those of  

the Pietra da Cantoni Group, with barnacle and coralline algae 

growing together, occur in the lower Miocene Latium-Abruzzi 

carbonate platform, where they have been interpreted as inner 

ramp deposits (Civitelli & Brandano 2005). Barnalgal assem-

blages are also reported in modern oceans, well below the fair 

weather wave base (Table 2; Macintyre & Milliman 1970; 

Milliman 1972). However, they do not represent an example 

of barnacles and coralline algae growing together but rather  

a case of barnacles colonizing a hard substrate provided by 

relict coralline algal bioconstructions (Milliman 1972). 

Although atypical barnacle facies are uncommon, they occur 

in both modern oceans and in the sedimentary record, thus 

indicating that dismissal of all barnacle concentrations as  

the result of nearshore carbonate factories might lead to erro-

neous palaeoenvironmental reconstructions. A perfect exam-

ple includes Bathylasmatidae-rich facies, which are dominated 

by barnacles while being related to bathyal settings 

(Buckeridge 1975, 1999; Newman & Ross 1976). The study of 

the barnacle assemblage, as demonstrated herein, can help  

in identifying these atypical situations. Furthermore, while  

the hard substrate where barnacle carbonate factories normally 

develop is a major site of skeletal production, due to the high 

hydraulic energy, few bioclasts are preserved there (Scoffin 

1988; Henrich et al. 1995). Most of the material is swept away 

and deposited in the closest sheltered areas (Scoffin 1988; 

Henrich et al. 1995). Remarkable transport processes are also 

reported in fossil barnacle facies, and can lead to significant 

displacement of the fossils (e.g., Buckeridge et al. 2018). 

Therefore, before making palaeoenvironmental assumptions 

based on barnacles, their preservation should be carefully 

evaluated in order to determine how far the material has been 

reworked and transported prior to burial.

Conclusions

Four barnacle facies from three different Burdigalian suc-

cessions have been analysed and divided into two groups on 

the basis of skeletal assemblages and barnacle preservation 

and diversity. The first group includes the barnamol facies of 

the Sandy Molasse Unit of the Sommières Basin (France) and 

the two barnacle facies of the Chilcatay Formation of the East 

Pisco Basin (Peru). These facies are overwhelmingly domi-

nated by shallow-water hard-substrate biota, including diffe-

rent species of barnacles, molluscs (mainly ostreids and 

pectinids) and echinoids; they are also characterized by  

an important siliciclastic fraction. Skeletal assemblages, 

Fig. 9. World map including the location of the fossil (Neogene and 

Quaternary) and recent/sub-recent barnacle facies. Circles represent 

recent/sub-recent facies; squares indicates fossil facies; stars identify 

the early Miocene facies investigated in this study.

background image

587

MIOCENE BARNACLE FACIES FROM EUROPE AND SOUTH AMERICA

GEOLOGICA CARPATHICA

, 2018, 69, 6, 573–592

barnacle preservation and sedimentary structures suggest that 

the bioclasts originated in a high-energy rocky shoreface envi-

ronment (less than 15–20 m of water depth), where reworking 

of bioclasts was significant. The second group includes  

the barnacle and coralline algae facies of the Pietra da Cantoni 

Group of the Tertiary Piedmont Basin (Italy). For this peculiar 

skeletal assemblage, rarely reported from the fossil record,  

the new name “barnalgal” is proposed. The barnalgal is 

charac terized by a single barnacle species, abundant coralline 

algae, benthic foraminifera and almost no siliciclastic ele-

ments. Compared to the other facies, barnacle shells are also 

better preserved. The related carbonate factory probably 

developed in slightly deeper conditions (20–40 m of water 

depth) than those responsible for the first group of barnacle 

facies. In this setting, the only hard substrate available for bar-

nacles were rhodoliths, leading to the formation of this atypical 

association. Reworking was also less important, reducing  

the abrasion of the specimens and leading to the conservation 

of abundant barnacle opercula in the sediment.

The analysis of both modern and fossil barnacle facies sug-

gests that the major factors controlling the development of  

a barnacle-dominated carbonate factory are the availability of 

Location

Age

Key 

references

Barnacle taxa

Associated bioclasts

Texture

Clastic 

particles

Palaeoenvironmental 

interpretation

Southeastern 

Japan, 

Shikoku

Oligocene 

to Miocene Sakai 1987

///

Very minor contributions from 

molluscs, bryozoans, echinoids, 

sponges and foraminifera

Coarse

Highly 

impure 

limestone

Shallow-water; high-energy

Austria, 

Bohemian 

Massif

Early 

Miocene

Nebelsick 

1989, 1992

///

Codominance with bivalves 

(pectinids and ostreids), important 

contribution from bryozoans and 

serpulids

Very coarse

50 %

Shallow-water; near-shore; high-

energy

Central Italy 

(several 

sites)

Early 

Miocene

Civitelli & 

Brandano 

2005

///

Molluscs, coralline algae, 

echinoids, bryozoans, serpulids, 

ostracods and benthic foraminifera 

(both large- and small-sized forms)

Very coarse

Almost pure 

limestone

Tropical; shallow-water inner 

ramp; hard substrate; high 

energy; high nutrient 

concentration

Motutapu 

Island, New 

Zealand

Early 

Miocene

Buckeridge 

1975

Hexelasma 

aucklandicum 

(Hector, 1888)

Bryozoans, molluscs, corals

Fine

Abundant

Deep-water; low-energy

Japan, 

Sendai

Early to 

middle 

Miocene

Nomura & 

Maeda 2008

Arossia sendaica 

(Hatai et al., 1976) 

(frequent); Balanus 

sulcatus Bruguière, 

1789 (frequent); B. 

crenatus Bruguière, 

1789; B. rostratus 

(Hoek, 1883)

Minor contributions from 

molluscs, serpulids and bryozoans; 

rare brachiopods, foraminifera, 

echinoids and corals

Very coarse

>>50 %

Very shallow-water; high-energy

California, 

Santa 

Monica 

Mountains

Early–

middle 

Miocene

Stanton & 

Alderson 

2013

///

Codominance with bivalves; minor 

contributions from echinoids, 

serpulids, bryozoans; rare 

gastropods and brachiopods; very 

rare coralline algae

Fine to 

coarse

Impure 

limestone

Warm-temperate; inner shelf; 

moderate-energy; hard substrate; 

high productivity

Italy, Liguria Middle 

Miocene

Brandano et 

al. 2015

///

Bryozoans, serpulids, echinoids, 

benthic foraminifera (including 

symbiont-bearing Amphistegina 

d’Orbigny, 1826), bivalves; minor 

contributions from Halimeda 

Lamouroux, 1812

Very coarse

From more 

than 50 % to 

around 20 %

Humid tropical; high terrigenous 

input; near-shore shallow-water, 

above fair-weather wave base; 

high energy; high nutrient supply

Spain, 

Almeria

Late 

Miocene

Doyle et al. 

1997

Megabalanus cf. 

tintinnabulum 

(Linnaeus, 1758)

Minor contributions from bivalves, 

echinoids and bryozoans

Very coarse

>50 %

Very shallow-water

Spain, 

Almeria

Late 

Miocene

Betzler et al. 

2000

///

Minor contributions from bivalves, 

bryozoans; coralline algae, 

echinoderms, gastropods; rare 

ostracods

Very coarse

///

More than 15m of water depth; 

moderate energy

Spain, 

Almeria

Pliocene

Aguirre et al. 

2008

Concavus concavus 

(Bronn, 1831) 

(dominant); 

Perforatus 

perforatus 

(Bruguière, 1789); 

Megabalanus Hoek, 

1913

Mainly bivalves but also bryozoans Coarse to 

very coarse

Up to 50 %

High energy; near-shore shallow-

water; high nutrient supply

Greece, 

Rafina

Pliocene

Radwańska & 

Radwański 

2008

Concavus concavus

///

///

///

///

New 

Zealand, 

North Island 

(several 

sites)

Pliocene–

Pleistocene

Hayton et al. 

1995; Kamp 

et al. 1988;

Buckeridge 

1983; 

Buckeridge 

2015; 

Buckeridge et 

al. 2018

Notobalanus 

Newman & Ross, 

1976; 

Austromegabalanus 

Newman, 1979; 

Balanus Costa, 

1778; Fosterella 

Buckeridge, 1983

Mainly bivalves, but also 

bryozoans; minor contributions 

from echinoids and small benthic 

foraminifera; rare coralline algae; 

brachiopods and solitary corals

Coarse to 

very coarse

10 % –20 % 

for most of 

the samples

Cold- to warm-temperate; high 

nutrient concentration; sub-tidal, 

less than 50–30 m of water 

depth; hard or coarse-grained 

substrate; high-energy seaway; 

locally important downslope 

transport of sediment

Table 3: Neogene and Quaternary fossil barnacle facies, including information on the position, age, barnacle assemblage, skeletal assemblage, 

texture, clastic fraction and palaeo-environmental interpretation. 

background image

588

COLETTI, BOSIO, COLLARETA, BUCKERIDGE, CONSANI and EL KATEB

GEOLOGICA CARPATHICA

, 2018, 69, 6, 573–592

hard substrate, hydrodynamic energy and nutrient availability. 

Consequently, the most favourable setting for a barnacle- 

dominated carbonate factory is probably an inshore rocky sub-

strate characterized by high-energy conditions and abundant 

nutrient supply. This situation can lead to the formation of  

the typical barnacle-rich deposit, where abundant barnacles are 

associated with other hard substrate organisms (e.g., the bar-

namol assemblage). The first group of facies recognized in 

this paper, embracing the French and Peruvian case studies, 

can be easily included in this category. The Italian barnalgal 

facies, on the other hand, represents an atypical barnacle facies 

related to situations that deviate from the aforementioned 

environmental optimum. Similar atypical situations can be 

identified throughout careful analyses of both the skeletal 

assemblage and the barnacle association, highlighting 

 

the importance of barnacle palaeontology for palaeoenviron-

mental reconstructions in shallow-water settings.

Acknowledgements:  The authors are grateful to Claudio  

Di Celma (Università degli Studi di Camerino) for his assis-

tance during the preparation of the manuscript. Special thanks 

go to Alfredo Frixa (Eni Spa), Jean Yves-Reynaud (Université 

de Lille) and Mario Urbina (Museo de Historia Natural de la 

Universidad Nacional Mayor de San Marcos) for their help 

during the field work in the Pietra da Cantoni area (Italy) and 

Sommières Basin (France), and East Pisco Basin (Peru), 

 respectively. Roberto Badano (Università di Genova) is grate-

fully acknowledged for his support with the XRD analyses. 

The authors acknowledge Silvia Spezzaferri (Université de 

Fribourg) for her help with the identification of foraminifera. 

The authors are also grateful to Daniela Basso and Elisa 

 Malinverno (both at Università di Milano Bicocca), Giovanni 

Bianucci (Università di Pisa) and Thomas J. DeVries (Univer-

sity of Washington) for their suggestions and fruitful discus-

sions. Comments and suggestions by Milan Kohút (Ústav vied 

o Zemi Slovenskej akadémie vied), Matúš Hyžný (Univerzita 

Komenského v Bratislave), Tomáš Kočí (Národní muzeum v 

Praze), Mathias Harzhauser (Naturhistorisches Museum 

Wien), and an anonymous reviewer greatly improved the qua-

lity of this work. Fieldwork by Giulia Bosio and Alberto 

 Collareta in the East Pisco Basin was supported by a grant 

from the Italian Ministero dell’Istruzione, dell’Università 

e della Ricerca [PRIN Project, 2012YJSBMK] to Giovanni 

Bianucci, Claudio Di Celma and Elisa Malinverno, and by  

a grant from the Università di Pisa to Giovanni Bianucci 

[PRA_2017_0032].

References

Aguirre  J.,  Martín  J.M.,  Braga  J.C.,  Betzler  C.,  Berning  B.  & 

 Buckeridge J.S. 2008: Densely packed concentration of sessile 

barnacles (Cirripedia: Sessilia) from the Early Pliocene of SE 

Spain. Facies 54, 2, 193–206.

Allen P., Mange-Rajetzky M, Matter A. & Homewood P. 1985: 

 Dynamic paleogeography of the open Burdigalian seaway. Swiss 

Molasse Basin. Eclogae Geol. Helv. 78, 2, 351–381.

Berger G.M. 1974: Carte géologique de la France à 1/50000, 

 Sommières  (n°964).  BRGM, Orléans.

Betzler C, Martin J.M. & Braga J.C. 2000: Non-tropical carbonates 

related to rocky submarine cliffs (Miocene, Almería, southern 

Spain). Sediment. Geol. 131, 51–65. 

Bianucci G., Landini W. & Buckeridge J. 2006: Whale barnacles and 

Neogene cetacean migration routes. New Zeal. J. Geol. Geop. 

49, 115–120.

Bianucci G., Urbina M. & Lambert O. 2015: A new record of Noto­

cetus vanbenedeni (Squalodelphinidae, Odontoceti, Cetacea) 

from the early Miocene of Peru. C. R. Palevol. 14, 1, 5–13.

Bianucci G., Di Celma C., Collareta A., Landini W., Post K.,  

Tinelli C., de Muizon C., Bosio G., Gariboldi K., Gioncada A., 

Malinverno E. & Cantalamessa G., Altamirano-Sierra. A, 

Salas-Gismondi R., Urbina M. & Lambert O. 2016a: Fossil 

 marine vertebrates of Cerro Los Quesos: Distribution of ceta-

ceans, seals, crocodiles, seabirds, sharks, and bony fish in  

a late Miocene locality of the Pisco Basin, Peru. J. Maps 12, 5, 

1037–1046.

Bianucci G., Di Celma C., Landini W., Post K., Tinelli C., de Muizon C., 

Gariboldi K., Malinverno E., Cantalamessa G., Gioncada A., 

Collareta A., Salas-Gismondi R., Varas-Malca R.M., Urbina M. 

& Lambert O. 2016b: Distribution of fossil marine vertebrates in 

Cerro Colorado, the type locality of the giant raptorial sperm 

whale Livyatan melvillei (Miocene, Pisco Formation, Peru).  

J. Maps 12, 3, 543–557.

Bianucci G., Collareta A., Bosio G., Landini W., Gariboldi K., 

 Gioncada A., Lambert O., Malinverno E., de Muizon C., Varas-

Malca R., Villa I.M, Coletti G., Urbina M. & Di Celma C. 2018: 

Taphonomy and palaeoecology of the lower Miocene marine 

vertebrate assemblage of Ullujaya (Chilcatay Formation, East 

Pisco Basin, southern Peru). Palaeogeogr. Palaeoclimatol. 

 Palaeoecol. 511, 256–279.

Bicchi E., Dela Pierre F., Ferrero E., Maia F., Negri A., Pirini- 

Radrizzani C., Radrizzani S. & Valleri G. 2006: Evolution of  

the Miocene carbonate shelf of Monferrato (North-western 

 Italy). Boll. Soc. Paleontol. It. 45, 2, 1–24. 

Brandano M., Tomassetti L. & Frezza V. 2015: Halimeda dominance 

in the coastal wedge of Pietra di Finale (Ligurian Alps, Italy): 

The role of trophic conditions. Sediment. Geol. 320, 30–37.

Brasier M.D. 1995a: Fossil indicators of nutrient levels 1: eutrophica-

tion and climate change. In: Bosence D.W.J. & Allison P.A. 

(Eds.): Marine paleoenvironmental analysis from fossils. Geol. 

Soc. London Spec. Publ. 83, 113–132.

Brasier M.D. 1995b: Fossil indicators of nutrient levels 2: evolution 

and extinction in relation to oligotrophy. In: Bosence D.W.J. & 

Allison P.A. (Eds.): Marine paleoenvironmental analysis from 

fossils. Geol. Soc. London Spec. Publ. 83, 133–150.

Bronn H.G. 1831: Italiens Tertiär-Gebilde und deren organische Ein-

schlüsse: vier Abhandlungen. Groos, Heidelberg, 1–176.

Bruguière J.G. 1789: Encyclopédie méthodique ou par ordre de 

matières. Histoire Naturelle des Vers, volume 1. Pancoucke

 Paris,  1–344.

Buckeridge J.S. 1975: The significance of cirripedes to the paleoeco-

logy of Motutapu Island. Tane 21, 121–129.

Buckeridge J.S. 1983: Fossil barnacles (Cirripedia: Thoracica) of 

New Zealand and Australia. New Zeal. Geol. Surv. Paleontol. 

Bull. 50, 1–151.

Buckeridge J.S. 1999: A new deep sea barnacle, Tetrachaelasma 

 tasmanicum sp. nov. (Cirripedia: Balanomorpha) from the South 

Tasman Rise, South Pacific Ocean. New Zeal. J. Mar. Fresh. 33, 

4, 521–531.

Buckeridge J.S. 2015: Revision of Southern Hemisphere taxa 

 

referred to Fosterella (Crustacea: Cirripedia), and their extinc-

tion in response to Pleistocene cooling. Integr. Zool. 10, 6,  

555–571.

background image

589

MIOCENE BARNACLE FACIES FROM EUROPE AND SOUTH AMERICA

GEOLOGICA CARPATHICA

, 2018, 69, 6, 573–592

Buckeridge J.S., Beu A. & Gordon D. 2018: Depositional environ-

ment of the early Pleistocene Castlepoint Formation, Wairarapa, 

New Zealand. New Zeal. J. Geol. Geop. 61, 4, 524–542.

Carannante G., Esteban M., Milliman J.D. & Simone L. 1988: 

 Carbonate lithofacies as paleolatitude indicators: problems and 

limitations. Sediment. Geol. 60, 333–346.

Carriol R.P. 2000: A new genus and a new species of Concavinae 

(Cirripedia, Thoracica) from the Serravallian of Mios (Gironde, 

France). N. Jb. Geol. Paläontol. Mh. 2, 122–128.

Carriol R.P. & Álvarez-Fernández E. 2015: Balanomorphs from late 

Upper Pleistocene and Holocene caves in northern Spain, with  

a new genus and species, and their palaeoclimatological implica-

tions. Ann. Paléontol. 101, 1, 21–27.

Carriol R.P. & Menkveld-Gfeller U. 2010: Balanoidea (Crustacea, 

Cirripedia) from the Upper Marine Molasse (Early Miocene) of 

the Bern area, Switzerland. Contrib. Nat. Hist. 13, 3–17.

Carriol R.P. & Schneider S. 2008: A new Concavinae (Cirripedia, 

Chesaconcavus) from the Late Burdigalian (Miocene) of Lower 

Bavaria (Germany). N. Jb. Geol. Paläontol. Abh. 248, 3, 345–354.

Carriol R.P., de Muizon C. & Secretan S. 1987: Les crustacés (Cirri-

pedia et Decapoda) du Néogène de la côte Péruvienne. Ann. 

Paléontol. 73, 137–164.

Chung F.H. 1974: Quantitative interpretation of X-ray diffraction 

 patterns of mixtures. I. Matrix-flushing method for quantitative 

multicomponent analysis. J. Appl. Cryst. 7, 6, 519–525.

Civitelli G. & Brandano M. 2005: The “Calcari a Briozoi e Lito-

tamni” in the Latium-Abruzzi carbonate platform (Central Apen-

nines, Italy): Atlas of the lithofacies and depositional model. 

[Atlante delle litofacies e modello deposizionale dei Calcari  

a Briozoi e Litotamni nella Piattaforma carbonatica laziale- 

abruzzese.] Boll. Soc. Geol. It. 124, 3, 611–643 (in Italian with 

 English  abstract).

Clari P., Dela Pierre F., Novaretti A. & Timpanelli M. 1995: Late Oli-

gocene–Miocene sedimentary evolution of the critical 

Alps-Apennines junction: the Monferrato area, Northwestern 

 Italy.  Terra Nova 7, 2, 144–152.

Coletti G., Basso D., Frixa A. & Corselli C. 2015: Transported rhodo-

liths witness the lost carbonate factory: a case history from  

the Miocene Pietra da Cantoni limestone (NW Italy). Riv. It. 

 Paleontol.  Strat.  121, 3, 345–368.

Coletti G., El Kateb A., Basso D., Cavallo A. & Spezzaferri S. 2017: 

Nutrient influence on fossil carbonate factories: Evidence from 

SEDEX extractions on Burdigalian limestones (Miocene, NW 

Italy and S France). Palaeogeogr. Palaeoclimatol. Palaeoecol. 

475, 80–92.

Coletti G., Basso D. & Corselli C. 2018: Coralline algae as depth 

 indicators in the Sommières Basin (early Miocene, Southern 

France). Geobios 51, 1, 15–30.

Collareta A., Bosselaers M. & Bianucci G. 2016a: Jumping from 

 turtles to whales: a Pliocene fossil record depicts an ancient dis-

persal of Chelonibia on mysticetes. Riv. Ital. Paleontol. Strat. 

122, 2, 35–44.

Collareta A., Margiotta S., Varola A., Catanzariti R., Bosselaers M. & 

Bianucci G. 2016b: A new whale barnacle from the early Pleis-

tocene of Italy suggests an ancient right whale breeding ground 

in the Mediterranean. C. R. Palevol 15, 5, 473–481.

Cornwall I.E. 1956: Identifying recent and fossil barnacles by the 

 figures in the shell. J. Paleontol. 30, 3, 646–561.

Cornwall I.E. 1958: Identifying recent and fossil barnacles. Can.  

J. Zool. 36, 1, 79–89.

Cornwall I.E. 1959: More shell figures and notes on barnacles. Can. 

J. Zool. 37, 4, 401–406.

Cornwall I.E. 1960: Barnacle shell figures and repairs. Can. J. Zool. 

38, 4, 827–832.

Cornwall I.E. 1962: The identification of barnacles with further 

 figures and notes. Can. J. Zool. 40, 4, 621–629.

Costa E.M. da 1778: Historia Naturalis Testaceorum Brittaniae [or 

The British Conchology, containing the natural history of the 

shells of Great Britain and Ireland]. The Author, London, 1–254.

D’Atri A., Dela Pierre F., Lanza R. & Ruffini R. 1999: Distinguishing 

primary and resedimented vitric volcanoclastic layers in the 

 Burdigalian carbonate shelf deposits in Monferrato (NW Italy). 

Sediment. Geol. 129, 143–163. 

D’Atri A., Dela Pierre F., Novaretti A., Cosca M.A. & Hunziker J.C. 

2001: Miocene plankton biostratigraphy and 

40

Ar/

39

Ar dating. 

Eclogae Geol. Helv. 94, 137–144.

d’Orbigny A.D. 1826: Tableau méthodique de la classe des Céphalo-

podes. Ann. Sci. Nat. 7, 96–169, 245–314.

Darwin C. 1854: A monograph of the sub-class Cirripedia, with 

 figures of all the species. The Balanidae, The Verrucidae, etc. 

Ray Society, London, 1–770.

Davadie C. 1963: Étude des balanes d’Europe et d’Afrique.  

Editions du Centre National de la Recherche Scientifique, Paris, 

1–146.

De Alessandri G. 1906: Monographical studies on the fossil cirri-

pedes of Italy. [Studi monografici sui cirripedi fossili d’Italia.] 

Palaeontographia Ital. 12, 207–324 (in Italian).

Demarcq G. 1970: Étude stratigraphique du Miocène rhodanien. 

Mem. BRGM 61, 1–257.

Dercourt J., Gaetani M., Vrielynk B., Barrier E., Biju-Duval B., 

 Brunet M.F., Cadet J.P., Crasquin S. & Sandulescu M. 2000: 

 Atlas Peri-Tethys, Palaeogeographical maps. Commission de la 

Carte Géologique du Monde, Paris, 1–279.

DeVries T.J. 1998: Oligocene deposition and Cenozoic sequence 

boundaries in the Pisco Basin (Peru). J. South Am. Earth Sci. 11, 

217–231.

DeVries T.J. & Frassinetti D. 2003: Range extensions and biogeo-

graphic implications of Chilean Neogene mollusks found in 

Peru. Bol. Mus. Nac. Hist. Nat. Chile 52, 119–135.

DeVries T.J. & Jud N.A. 2018: Lithofacies Patterns and Paleogeo-

graphy of the Miocene Chilcatay and lower Pisco Depositional 

Sequences (East Pisco Basin, Peru). Bol. Soc. Geol. Perú, volu-

men jubilar 8, 124–167.

DeVries T.J. & Schrader H. 1997: Middle Miocene marine sediments 

in the Pisco Basin (Peru). Bol. Soc. Geol. Perú 87, 1–13.

Di Celma C., Malinverno E., Bosio G., Collareta A., Gariboldi K., 

Gioncada A., Molli G., Basso D., Varas-Malca R.M., 

 

Pierantoni P.P., Villa I.M., Lambert O., Landini W., Sarti G., 

Cantalamessa G., Urbina M. & Bianucci G. 2017: Sequence stra-

tigraphy and paleontology of the upper Miocene Pisco Forma-

tion along the western side of the lower Ica valley (Ica Desert, 

Peru). Riv. It. Paleontol. Strat. 123, 2, 255–274.

Di Celma C., Malinverno E., Bosio G., Gariboldi K., Collareta A., 

Gioncada A., Landini W., Pierantoni P.P. & Bianucci, G. 2018a: 

Intraformational unconformities as a record of late Miocene 

 eustatic falls of sea level in the Pisco Formation (southern Peru). 

J. Maps 14, 2, 

607–619.

Di Celma C., Malinverno E., Collareta A., Bosio G., Gariboldi K., 

Lambert O., Landini W., Gioncada A., Villa I.M., Coletti G.,  

de Muizon C., Urbina M. & Bianucci G. 2018b: Facies analysis, 

stratigraphy and marine vertebrate assemblage of the early 

 Miocene Chilcatay Formation at Ullujaya (Pisco basin, Peru).  

J. Maps 14, 2, 257–268.

Domack E.W. 1988: Biogenic facies in the Antartic glacimarine envi-

ronment: basis for a polar glacimarine summary. Palaeogeogr. 

Palaeoclimatol. Palaeoecol. 63, 357–372.

Dominici S., Bartalini M., Benvenuti M. & Balestra B. 2011: Large 

kings with small crowns: a Mediterranean Pleistocene whale 

barnacle. Boll. Soc. Paleontol. Ital. 50, 95–101.

Donovan S.K. 1988: Palaeoecology and taphonomy of barnacles 

from the Plio-Pleistocene Red Crag of East Anglia. Proc. Geol. 

Ass. 99, 4, 279–289.

background image

590

COLETTI, BOSIO, COLLARETA, BUCKERIDGE, CONSANI and EL KATEB

GEOLOGICA CARPATHICA

, 2018, 69, 6, 573–592

Dorreen J.M. 1948: A Foraminiferal fauna from the Kaiatan Stage 

(Upper Eocene) of New Zealand. J. Paleontol. 22, 3, 281–300.

Douvillé H. 1911: Les foraminifères dans le Tertiaire des Philippines. 

Philipp. J. Sci. 6, 2, 53-80.

Doyle P., Mather A.E., Bennet M.R. & Bussel M.A. 1997: Miocene 

barnacle assemblages from Southern Spain and their palaeoenvi-

ronmental significance. Lethaia 29, 3, 267–274.

Dunbar R.B., Marty R.C. & Baker P.A. 1990: Cenozoic marine sedi-

mentation in the Sechura and Pisco basins, Peru. Palaeogeogr. 

Palaeoclimatol. Palaeoecol. 77, 235–261.

Farrow G.E., Cucci M. & Scoffin T.P. 1978: Calcareous sediments on 

nearshore continental shelf of western Scotland. Proc. Roy. Soc. 

Edinburgh 76B, 55–76.

Flügel E. 2010: Microfacies of Carbonate Rocks: Analysis, Inter-

pretation and Application. 2

nd

 Edition. Springer, New York, 

1–1007.

Foster B.A. 1987: Barnacle ecology and adaptation. In: Southward 

A.J. (Ed.): Barnacle Biology. Crustacean Issues vol. 5. Balkema

Rotterdam, 113–133.

Foster B.A. & Buckeridge J.S. 1987: Barnacle paleontology. In: 

Southward A.J. (Ed.): Barnacle Biology. Crustacean Issues  

vol. 5. Balkema, Rotterdam, 43–62.

Frank T.D., James N.P., Bone Y., Malcolm I. & Bobak L.E. 2014: Late 

Quaternary carbonate deposition at the bottom of the world. 

 Sediment.  Geol.  305, 1–16.

Galloway J.J. 1933: A manual of Foraminifera. Principia Press

Bloomington, 1–483.

Gioncada A., Gariboldi K., Collareta A., Di Celma C., Bosio G., 

 Malinverno E., Lambert O., Pike J., Urbina M. & Bianucci G. 

2018: Looking for the key to preservation of fossil marine verte-

brates in the Pisco Formation of Peru: New insight from a small 

dolphin skeleton. Andean Geol. 45, 379–398.

Glynn P.W. & Wellington G.M. 1983: Corals and coral reefs of the 

Galapagos islands. University of California Press, Berkeley and 

Los Angeles, 1–330.

Gould A.A. 1841: Crustacea. In: Report on the Invertebrata of Massa-

chusetts, comprising the Mollusca, Crustacea, Annelida, and 

 Radiata. Folsom, Wells, and Thurston, Cambridge, 321–341.

Halfar J., Strasser M., Riegl B. & Godinez-Orta L. 2006: Oceano-

graphy, sedimentology and acoustic mapping of a bryomol 

 carbonate factory in the northern Gulf of California, Mexico. In: 

Pedley H.M. & Carannante G. (Eds.): Cool-water carbonates: 

depositional systems and palaeoenvironmental controls. Geol. 

Soc. London Spec. Publ. 255, 197–215.

Harzhauser M. & Piller W.E. 2007: Benchmark data of a changing 

sea: Palaeogeography, Palaeobiogeography and events in the 

Central Paratethys during the Miocene. Palaeogeogr. Palaeo­

climatol. Palaeoecol. 253, 831.

Harzhauser M., Newman W.A. & Grunert P. 2011: A new Early Mio-

cene barnacle lineage and the roots of sea-turtle fouling Cheloni-

biidae (Cirripedia, Balanomorpha). J. Systematic Palaeontol. 9, 

473–480.

Hatai K., Masuda K. & Noda H. 1973: Marine fossils from the 

 Moniwa Formation distributed along the Natori River, Sendai, 

Northeast Honshu Japan. Part 1. Brachiopoda from the Moniwa 

Formation. Saito Ho­on Kai Mus. Res. Bull. 42, 9–19.

Hayton S., Nelson C.S. & Hood S.D. 1995: A skeletal assemblage 

classification system for non-tropical carbonate deposits based 

on New Zealand Cenozoic limestones. Sediment. Geol. 100, 

123–141.

Hector J. 1888: Specimens of a large fossil stalked cirripede. Trans. 

New Zeal. Inst. 20, 440.

Henrich L., Freiwald A., Betzler C., Bader B., Samtleben C., Brachert 

T.C., Wehrmann A., Zankl H. & Kühlmann D.H.H. 1995: 

 Controls on modern carbonate sedimentation on warm-tempe-

rate to Arctic coasts, shelves and seamounts in the Northern 

Hemisphere: implications for fossil counterparts. Facies 32, 1, 

71–108.

Hernández-Molina F.J., Fernández-Salas L.M., Lobo F., Somoza L., 

Díaz-del-Río V. & Alveirinho Dias J.M. 2000: The infralittoral 

prograding wedge: a new large-scale progradational sedimentary 

body in shallow marine environments. Geo­Mar. Lett., 20, 2, 

109–117.

Hollingworth N.T.J. & Barker M.J. 1991: Colour pattern preservation 

in the fossil record: taphonomy and diagenetic significance. In: 

Donovan S.K. (Ed.): The processes of fossilization. Belhaven 

Press, London, 105–119.

Hoek P.P.C. 1883: Report on the Cirripedia collected by H.M.S. Chal-

lenger during the years 1873–76. Zoology 8, 25, 1–169.

Hoek P.P.C. 1913: The Cirripedia of the Siboga-Expedition. B. Cirri-

piedia Sessilia. Siboga­Expeditie 31B, 129–275.

Hofker J. 1951: The toothplate-foraminifera. Arch. Néerland. Zool. 8, 

1, 353–373.

Hoskin C.M. & Nelson R.V. 1969: Modern marine carbonate sedi-

ment, Alexander Archipelago, Alaska. J. Sediment. Petrol. 39, 2, 

581–590.

Hottinger L.K. 1983: Neritic Macroid Genesis: an Ecological 

 Approach. In: Peryt T.M. (Ed.): Coated Grains. Springer­Verlag

Berlin, 38–55.

Kamp P.J.J., Harmsen F.J., Nelson C.S. & Boyle S.F. 1988: Barna-

cle-dominated limestone with giant cross-beds in a non-tropical, 

tide-swept, Pliocene forearc seaway, Hawke’s Bay, New Zealand. 

Sediment. Geol. 60, 173–195.

Keij A.J. 1969: Bicornifera lindenbergi n. sp. from the Upper Oligo-

cene of Escornebéou, S.W. France. N. Jb. Geol. Paläontol.  

Mh. 1, 241–246.

Klicpera A., Taylor P.D. & Westphal H. 2013: Bryoliths constructed 

by bryozoans in symbiotic associations with hermit crabs in  

a tropical heterozoan carbonate system, Golfe d’Arguin, Mauri-

tania. Mar. Biodiv. 43, 4, 429–444.

Kočí T., Veselská M.K., Newman W.A., Buckeridge J.S. & Sklenář J. 

2017: Archaeochionelasmus nekvasilovae gen. et sp. nov. (Cirri-

pedia, Balanomorpha, Chionelasmatoidea) from the Bohemian 

Cretaceous Basin (Czech Republic): the first bona fide Creta-

ceous neobalanoform. Zootaxa 4294, 2, 181–196. 

Lambert O., Bianucci G. & Urbina M. 2014: Huaridelphis raimondii

a new early Miocene Squalodelphinidae (Cetacea, Odontoceti) 

from the Chilcatay Formation, Peru. J. Vert. Paleontol. 34, 5, 

987–1004.

Lambert O., de Muizon C. & Bianucci G. 2015. A new archaic homo-

dont toothed cetacean (Mammalia, Cetacea, Odontoceti) from 

the early Miocene of Peru. Geodiversitas 37, 1, 79–108.

Lambert O., Bianucci G., Urbina M., Geisler J.H. 2017a: A new inioid 

(Cetacea, Odontoceti, Delphinidae) from the Miocene of Peru 

and the origin of modern dolphin and porpoise families. Zool. J. 

Linn. Soc. 179, 4, 919–946.

Lambert O., de Muizon C., Malinverno E., Di Celma C., Urbina M. & 

Bianucci G. 2017b: A new odontocete (toothed cetacean) from 

the Early Miocene of Peru expands the morphological disparity 

of extinct heterodont dolphins. J. Syst. Palaeontol. 16, 12, 

 981–1016.

Lamouroux J.V. 1812: Extrait d’un mémoire sur la classification des 

Polypiers coralligènes non entièrement pierreux. Nouv. Bull. Sci. 

Soc. Philom. Paris 3, 181–188.

Landini W., Altamirano-Sierra A., Collareta A., Di Celma C., Urbina 

M. & Bianucci G. 2017a: The late Miocene elasmobranch as-

semblage from Cerro Colorado (Pisco Formation, Peru). J. S. 

Am. Earth Sci. 73, 168–190.

Landini W., Collareta A., Pesci F., Di Celma C., Urbina M. &   

Bianucci G. 2017b: A secondary nursery area for the copper 

shark Carcharhinus brachyurus from the late Miocene of Peru. 

J. S. Am. Earth Sci. 78, 164–174.

background image

591

MIOCENE BARNACLE FACIES FROM EUROPE AND SOUTH AMERICA

GEOLOGICA CARPATHICA

, 2018, 69, 6, 573–592

Landini W., Collareta A., Di Celma C., Malinverno E., Urbina M. & 

Bianucci, G. 2018: The early Miocene elasmobranch assem-

blage from Zamaca (Chilcatay Formation, Peru). J. S. Am. Earth 

Sci. in press. doi:10.1016/j.jsames.2018.08.004.

León W., Aleman A., Torres V., Rosell W. & De La Cruz O. 2008: 

Stratigraphy, sedimentology and tectonic evolution of the Eastern 

Pisco basin. [Estratigrafía, sedimentología y evolución tectónica 

de la cuenca Pisco Oriental.] Bol. INGEMMET (Serie D), 27, 

1–144 (in Spanish).

Linnaeus C. 1758: Systema Naturae sive Regna Tria Naturae, secun-

dum Classes, Ordines, Genera, Species, cum characteribus, dif-

ferentiis, synonymis, locis, Tomus I. Editio decima, reformata. 

L. Salvius, Stockholm, 1–824 (in Latin).

Linnaeus C. 1767: Systema Naturae sive Regna Tria Naturae, secun-

dum Classes, Ordines, Genera, Species, cum characteribus, dif-

ferentiis, synonymis, locis, Tomus I. Editio duodecima, reforma-

ta. L. Salvius, Stockholm, 1–1327 (in Latin).

MacIntyre I.G & Milliman J.D. 1970: Physiographic features of the 

Outer Shelf and Upper Slope, Atlantic Continental Margin, 

Southeastern United States. Geol. Soc. Am. Bul. 81, 2577–2598. 

Maffione M., Speranza F., Faccenna C., Cascella A., Vignaroli G. & 

Sagnotti L. 2008: A syncronous Alpine and Corsica-Sardinia 

 rotation.  J. Geophys. Res. 113, doi:10.1029/2007JB005214.

Marx F.G., Collareta A., Gioncada A., Post K., Lambert O.,  Bonaccorsi 

E., Urbina M., Bianucci G. 2017: How whales used to filter: 

 exceptionally preserved baleen in a Miocene cetotheriid. J. Anat. 

231, 2, 212–220.

Massari F. & Chiocci F. 2006: Biocalcarenite and mixed cool-water 

prograding bodies of the Mediterranean Pliocene and Pleisto-

cene: architecture, depositional setting and forcing factors. In: 

Pedley H.M. & Carannante G. (Eds.): Cool-water carbonates: 

depositional systems and palaeoenvironmental controls. Geol. 

Soc. London Spec. Publ. 255, 95–120.

Massari F. & D’Alessandro A. 2012: Facies partitioning and sequence 

stratigraphy of a mixed siliciclastic carbonate ramp stack in the 

Gelasian of Sicily (S Italy): a potential model for icehouse, 

 distally-steepened heterozoan ramps. Riv. It. Paleontol. Strat. 

118, 3, 503–534.

Michel J., Mateu-Vicens G. & Westphal H. 2011: Modern Heterozoan 

carbonates from a eutrophic tropical shelf (Mauritania). 

 

J. Sediment. Res. 81, 9, 641–655.

Milliman J.D. 1972: Atlantic Continental Shelf and Slope of the United 

States-Petrology of the Sand Fraction of Sediments, Northern 

New Jersey to Southern Florida. US. Geol. Surv. Prof. Pap.  

529-J, 1–48.

Montfort P.D. de 1808: Conchyliologie systématique et classification 

méthodique des coquilles, Vol. 1. Schoell, Paris, 1–409.

Müller J. & Milliman J.D. 1973:  Relict carbonate-rich sediments on 

Southwestern Grand Bank, Newfoundland. Can. J. Earth Sci. 

10, 12, 1744–1750.

Murray J. 2006: Ecology and Applications of Benthic Foraminifera. 

Cambridge University Press, New York, 1–426.

Nebelsick J.H. 1989: Temperate Water Carbonate Facies of the Early 

Miocene Paratethys (Zogelsdorf Formation, Lower Austria). 

 Facies 21, 11–40.

Nebelsick J.H. 1992: Components analysis of sediment composition 

in Early Miocene temperate carbonates from the Austrian 

 Paratethys.  Palaeogeogr. Palaeoclimatol. Palaeoecol. 91,  

59–69.

Nelson C.S., Keane S.L. & Head P.S. 1988: Non-tropical carbonate 

deposits on the modern New Zealand shelf. Sediment. Geol. 60, 

71–94.

Newman W.A. 1979: On the biogeography of balanomorph barnacles 

of the Southern Ocean, including two new balanid taxa; a sub-

family, two genera and three species. In: Proceedings of the 

 International Symposium on Marine Biogeography and Evolution 

in the Southern Hemisphere. New Zealand Department of Scien­

tific and Industrial Research Information Series, Auckland, 137, 

279–306.

Newman W.A. 1982: A review of extant taxa of the “group of Balanus 

concavus” (Cirripedia: Thoracica) and a proposal for genus- 

group ranks. Crustaceana 43, 1, 25–36. 

Newman W.A. 1991: Origins of Southern Hemisphere endemism, 

 especially amongmarine Crustacea. Mem. Queensland Mus. 31, 

51–76.

Newman W.A. 1996: Sous-classe des Cirripèdes (Cirripedia Burmeis-

ter, 1834), superordre des Thoraciques et des Acrothoraciques 

(Thoracica Darwin, 1854–Acrothoracica Gruvel, 1905). In: 

 Forest J. (Ed.): Traité de Zoologie. Anatomie, Systématique, 

 Biologie. Crustacés, 7. Masson, Paris, 453–540.

Newman W.A. & Abbot D.P. 1980: Cirripedia: The Barnacles. In: 

Morris R.H., Abbot D.P. & Haderlie E.C. (Eds.): Intertidal inver-

tebrates of California. Stanford University Press, Stanford,  

504–535.

Newman W.A. & Ross A. 1971: Antarctic Cirripedia: monographic 

account based on specimens collected chiefly under the United 

States Antarctic Research Program, 1962–1965. Am. Geophys. 

Union. Antarct. Res. Ser. 14, 1–257.

Newman W.A. & Ross A. 1976: Revision of the balanomorph barna-

cles; including a catalogue of the species. Mem. San Diego Soc. 

Nat. Hist. 9, 1–108.

Newman W.A., Zullo V.A. & Wainwright S.A. 1967: A critique on 

recent concepts of growth in Balanomorpha (Cirripedia, 

 

Thoracica). Crustaceana 12, 2, 167–178.

Newman W.A., Zullo V.A. & Withers T.H. 1969: Cirripedia. In: 

Moore R.C. (Ed.): Treatise on Invertebrate Paleontology, Part R, 

Arthropoda 4, Vol. 1. University of Kansas and the Geological 

Society of America, Boulder, R206–295.

Nielsen J.K. & Funder S. 2003: Taphonomy of Eemian marine mol-

lusks and acorn barnacles from eastern Arkhangelsk region, 

northern Russia. Palaeogeogr. Palaeoclimatol. Palaeoecol. 191, 

139–168.

Nomura S. & Maeda H. 2008: Significance of autochtonous fossil 

barnacles from the Miocene Natori Group at the Moniwa-Goishi 

area, northeast Japan. Paleontol. Res. 12, 1, 63–79.

Novaretti A., Bicchi E., Condello A., Ferrero E., Maia F., Tonon M. & 

Torta D. 1995: The Oligo-Miocene succession from the 

 Mon ferrato area (NW Italy): a synthesis of biostratigraphic  

data. [La successione oligo-miocenica del Monferrato: sintesi 

dei dati biostratigrafici.] Atti del Convegno “Rapporti tra 

 

Alpi e Appennino”, Peveragno (CN). Acc. Naz. Sci., Collana 

Scritti e Documenti XIV, 39–59 (in Italian with English 

 abstract). 

Pilsbry H.A. 1916: The sessile barnacles (Cirripedia) contained in  

the collections of the U.S. National Museum, including 

 

a monograph of the American Species. Bull. U.S. Nat. Mus. 93, 

1–366.

Radwańska  U.  &  Radwański  A.  2008:  Eco-taphonomy  of  mass- 

aggregated giant balanids Concavus (Concavus)  concavus  

(Darwin, 1854) from the Lower Pliocene (Zanclean) of 

 

Rafina near Pikermi (Attica, Greece). Acta Geol. Pol. 58, 1, 

87–103.

Raymond P.E. & Stetson H.C. 1932: A calcareous beach on the coast 

of Maine. J. Sediment. Petrol. 2, 2, 51–62.

Reijmer J.J.G., Bauch T. & Schäfer P. 2012: Carbonate facies patterns 

in surface sediments of upwelling and non-upwelling shelf  

environments (Panama, East Pacific). Sedimentology 59, 1,  

32–56.

Reymond C.E., Zihrul K.S., Halfar J., Riegl B., Humphreys A., 

 Hildegard W. 2016: Heterozoan carbonates from the equatorial 

rocky reefs of the Galapagos Archipelago. Sedimentology 63, 4, 

940–958.

background image

592

COLETTI, BOSIO, COLLARETA, BUCKERIDGE, CONSANI and EL KATEB

GEOLOGICA CARPATHICA

, 2018, 69, 6, 573–592

Reynaud J.Y. & James N.P. 2012: The Miocene Sommières basin, SE 

France: Bioclastic carbonates in a tide-dominated depositional 

system. Sediment. Geol. 282, 360–373.

Rögl F. 1998: Palaeogeographic considerations for Mediterranean 

and Paratethys seaways (Oligocene to Miocene). Ann. Nat. Hist. 

Mus. Wien 99, 279–310.

Ross A. & Newman W.A. 1967: Eocene Balanidae of Florida, inclu-

ding a new genus and species with a unique plan of “turtle- 

barnacle” organization. Am. Mus. Nov. 2288, 1–21.

Rossi M., Mosca P., Polino R., Rogledi S. & Biffi U. 2009: New out-

crop and subsurface data in the Tertiary Piedmont Basin: uncon-

formity-bounded stratigraphic units and their relationships with 

basin-modification phases. Riv. Ital. Paleontol. Strat. 115, 3, 305–335.

Ruffini R. 1995: Evidence of tertiary volcanic activity in Western 

Alps: problems and hypotheses. [Evidenze di attività vulcanica 

terziaria nelle Alpi occidentali: problemi ed ipotesi.] PhD thesis, 

Torino, 1–160 (in Italian).

Sacco F. 1893: Sur quelques Tinoporinae du Miocène de Turin. Bull. 

Soc. Belge Géol., Paléont. et Hydrol. 7, 204–207.

Sakai H. 1987: Storm barnacles beds and their deformation in the 

Muroto-Misaki olistostrome and mélange complex, Shikoku.  

J. Geol. Soc. Jap. 83, 8, 617–620.

Sanford E. & Menge B.A. 2001: Spatial and temporal variation in 

barnacle growth in coastal upwelling system. Mar. Ecol. Progr. 

Ser. 209, 143–157.

Santos A., Mayoral E. & Muñiz F. 2005: Bioerosion scars of acorn 

barnacles from the southwestern Iberian Peninsula, upper Neo-

gene. Riv. It. Paleontol. Strat. 111, 1, 181–189.

Scarff J.E. 1986: Occurrence of the barnacles Coronula diadema,  

C. reginae and Cetopirus complanatus (Cirripedia) on right 

whales. Sci. Rep. Whal. Res. Inst. 37, 129–153.

Scoffin T.P. 1988: The environments of production and deposition of 

calcareous sediments on the shelf west of Scotland. Sediment. 

Geol. 60, 107–124.

Scoffin T.P. & Bowes G.E. 1988: The facies distribution of carbonate 

sediments on Porcupine Bank, northeast Atlantic. Sediment. 

Geol. 60, 125-134.

Schüttenhelm R.T.E. 1976: History and modes of Miocene carbonate 

deposition in the interior of the Piedmont Basin, NW Italy. 

Utrecth Micropaleontol. Bull. 4, 1–207.

Seilacher A. 2005: Whale barnacles: exaptational access to a forbid-

den paradise. Paleobiology 31 (Suppl.), 27–35.

Stanton R.J. & Alderson J.M. 2013: Limestone interbedded with sub-

marine volcanics: the Early-Middle Miocene Conejo Volcanics, 

California. Facies 59, 3, 467–480.

Taviani M., Reid D.E. & Anderson J.B. 1993: Skeletal and isotopic 

composition and paleoclimatic significance of late Pleistocene 

carbonates, Ross Sea, Antartica. J. Sediment. Petrol. 63, 1,  

84–90.

Vannucci G., Piazza M., Fravega P. & Abate C. 1996: Lithostrati-

graphy and paleoecology of the rhodolithic lithosomes of the 

“Pietra da Cantoni” (Eastern Monferrato, NW Italy). [Lito-

stratigrafia e paleoecologia di successioni a rodoliti della Pietra 

da Cantoni (Monferrato Orientale, Italia Nord-Occidentale).] 

Atti Soc. Tosc. Sci. Nat. Mem. Ser. A 103, 69–86 (in Italian with 

English abstract).

Westphal H., Halfar J. & Freiwald A. 2010: Heterozoan carbonates in 

subtropical to tropical settings in the present and past. Int. J. 

Earth Sci. 99, S1, S153–S169.

Wilson J.B. 1988: A model for temporal changes in the faunal 

 composition of shell gravels during a transgression on the 

 continental shelf around the British Isles. Sediment. Geol. 60, 

95–105.

Wright R., Dunbar R.B., Allen M. & Baker P. 1988: Morphology of 

stacked marine delta lobes, East Pisco Basin, Peru. In: Bally, 

A.W. (Ed.), Atlas of seismic stratigraphy. AAPG Studies in 

 Geology 27, 192

196.

Zullo V.A. 1984: New genera and species of balanoid barnacles from 

the Oligocene and Miocene of North Carolina. J. Paleontol. 58, 

5, 1312

1338.

Zullo V.A. 1992: Revision of the balanid barnacle genus Con cavus 

Newman, 1982, with the description of a new subfamily,  

two genera and eight new species. J. Paleontol. 66, 6 

 

(Suppl.), 1

46.