Sei sulla pagina 1di 19

Sedimentary Geology 335 (2016) 197–215

Contents lists available at ScienceDirect

Sedimentary Geology

journal homepage: www.elsevier.com/locate/sedgeo

Determining flow directions in turbidites: An integrated


sedimentological and magnetic fabric study of the Miocene Marnoso
Arenacea Formation (northern Apennines, Italy)
Fabrizio Felletti ⁎, Eleonora Dall'Olio, Giovanni Muttoni
Università degli Studi di Milano, Dipartimento di Scienze della Terra ‘Ardito Desio’, Via Mangiagalli 34, I-20133, Milano, Italy

a r t i c l e i n f o a b s t r a c t

Article history: Depositional models of turbidity flows require estimating paleocurrent directions using sedimentological indica-
Received 20 November 2015 tors such as flute and ripple marks, but these are not always present in outcrop sections or drill cores. In this
Received in revised form 5 February 2016 study, we apply the anisotropy of magnetic susceptibility (AMS) as an alternative tool to estimate paleocurrent
Accepted 7 February 2016
directions in a case-study turbiditic system: the Marnoso Arenacea foredeep turbidites of Miocene age exposed
Available online 17 February 2016
in the northern Apennines of Italy. Different depositional facies have been sampled for AMS and additional
Editor: Dr. B. Jones rock-magnetic analyses. We observed a good agreement between paleocurrent directions from flute casts at
the base of turbidite beds and mean directions of maximum magnetic susceptibility axes in organized facies
Keywords: such as massive and laminated sandstones, even if a relatively small but apparently consistent offset of
Turbidites ~15–20° seems to be present. Highly dispersed AMS fabrics were instead observed in disordered facies such as
Marnoso Arenacea Formation convoluted and undulated sandstones as well as debrites. This strong correlation between hydrodynamic re-
Miocene gimes of depositional facies and AMS data represents a novel contribution and confirms the validity of the
Anisotropy of magnetic susceptibility AMS method to estimate flow directions in the absence of sedimentological indicators. Finally, paleomagnetic
Paleocurrent directions
analyses from the literature were used to reconstruct the paleogeography of the Marnoso Arenacea basin and
make inferences about the origin and direction of transport of the sediments at the basin scale.
© 2016 Elsevier B.V. All rights reserved.

1. Introduction However, sedimentological indicators are not always present in outcrop


sections and are virtually absent from drill cores, and this latter consti-
Submarine density flows characterize sediment transport in vast tutes a major limitation when it comes to reconstructing, for example,
portions of the deep sea, and produce some of the most extensive and the depositional architecture of turbidites from Atlantic-type passive
voluminous sediment accumulations on Earth that can contain major margin basins that have proven to be among the most important set-
oil and gas reserves (Viana, 2008; Nilsen et al., 2007). Recent studies tings for modern oil exploration.
have shown that turbiditic systems are characterized by great variability An alternative method to estimate paleocurrents is represented by
in terms of size, geometry, facies, and stacking patterns, as a result of the the analysis of the anisotropy of magnetic susceptibility (AMS), which
interplay of several factors such as sea-level change, local and regional is a relatively inexpensive and highly effective technique that investi-
tectonic settings, basin size and shape, sediment type and frequency of gates the magnetic fabric of rocks as a proxy of grains spatial distribu-
depositional events, and volume of gravity flows (Mutti, 1992; tion. Standing water deposition is characterized by planar, gravity-
Pickering et al., 1995; Mulder and Alexander, 2001; Carruba et al., induced magnetic settling fabrics with minimum susceptibility axes of
2004; Nilsen et al., 2007; Mutti et al., 2009; Felletti and Bersezio, magnetic grains oriented perpendicular to the depositional plane within
2010; Talling et al., 2015, Maino et al., 2013). which maximum and intermediate susceptibility axes are uniformly
The evaluation of paleocurrent directions is a basic requirement to dispersed, whereas the magnetic fabric of sediments deposited under
interpret the depositional architecture of turbiditic successions as well the action of flowing water is typified by a current-oriented magnetic
as the structure and texture of the constituent beds. Paleocurrents are foliation plane (e.g., Ellwood, 1980; Lowrie and Hirt, 1987; Taira,
traditionally estimated through sedimentological indicators (i.e., ripple 1989; Parés et al., 2007; Schwehr et al., 2007; Novak et al., 2014). Syn-
marks, flute marks) or, more rarely, by means of optical analyses on depositional or early post-depositional soft-sediment deformation and
thin sections aimed at determining the orientation of elongated grains. the modifying effect of burrowing organisms are usually considered as
shaping processes of primary fabrics, although it is important to realize
⁎ Corresponding author. that these processes may significantly alter the original fabric formed by
E-mail address: fabrizio.felletti@unimi.it (F. Felletti). the depositing current (Hiscott and Middleton, 1980). Also, secondary

http://dx.doi.org/10.1016/j.sedgeo.2016.02.009
0037-0738/© 2016 Elsevier B.V. All rights reserved.
198 F. Felletti et al. / Sedimentary Geology 335 (2016) 197–215

fabrics due to post-depositional compaction, downslope creep, diagen- 2015) that crops out extensively over an area 180 km long and 40 km
esis, or even metamorphism may overprint or even completely obliter- wide in the northern Apennines from Emilia to Umbria (Fig. 1A), and
ate the primary current-induced fabric (Allen, 1984; Schwehr et al., is partly buried under tectonic or sedimentary units both to the W and
2007; Novak et al., 2014). to the NE of its present outcrop area. The Marnoso Arenacea Formation
In this study, we apply the AMS to estimate paleoflow directions in is a composite, wedge-shaped turbidite succession representing the
the Marnoso Arenacea Formation, which is a turbiditic system of Mio- final filling of a Miocene migrating Apennine foredeep complex, which
cene age extensively outcropping in the northern Apennines of Italy. accumulated between the Langhian and the Tortonian (Ricci Lucchi
Our contribution has two main targets: and Valmori, 1980; Fig. 1B). Palinspastic restorations indicate an origi-
nal width of the Marnoso Arenacea Formation basin of 140–90 km
1. Validating the AMS-based paleocurrent estimates. The well-exposed
and a sediment thickness of up to 3 km (Ricci Lucchi, 1975; Boccaletti
stratigraphic sections selected in this study are characterized by the
et al., 1990; Vai, 2001; Costa et al., 1998). The 1 km-isopach (Dondi
presence of clear sedimentological indicators of paleocurrent direc-
et al., 1982; Argnani and Ricci Lucchi, 2001) defines an elongate basin
tions (flute casts and ripple marks) that can be used for checking
extending for at least 400 km along the Apennine front (Fig. 1A). To
the reliability of paleocurrent estimates from AMS data.
the southeast, the thickness of the Marnoso Arenacea Formation de-
2. Determining which type of sediment works best for the application
creases rapidly, suggesting the presence of a structural high within the
of the AMS method. We tackled this target by sampling 11 different
southeastern (Umbrian) part of the basin (Argnani and Ricci Lucchi,
depositional intervals (e.g., including massive, laminated, and convo-
2001). Paleogeographic reconstructions as well as petrographic and
luted sandstones, debrites, as well as white marlstone beds separat-
paleocurrent analyses revealed longitudinal transport patterns within
ing turbiditic intervals).
the elongated Marnoso Arenacea Formation basin, fed mainly by Alpine
The main outcome of this analysis is a systematic study of the rela- (crystalline) sources through multiple entry points located to the N and
tionships between AMS and sedimentary structures that brings the ap- W in present-day coordinates. Most paleocurrent directions measured
plication of AMS techniques in turbidites to new levels of at the base of turbidite beds indicate axial flows from the NW to the
understanding. SE (in present-day coordinates; Fig. 1A). In addition, minor volumes of
carbonate and hybrid turbidites, derived from shallow-water carbonate
2. Geological setting platforms located along the southern and southeastern margins of the
basin (Fig. 1A), have also been reported to display flow directions
The Marnoso Arenacea Formation is a non-channelized, mainly from the SE to the NW, i.e. opposite to the siliciclastic turbidites
siliciclastic turbidite system (Ricci Lucchi and Valmori, 1980; Amy and (Gandolfi et al., 1983). Furthermore, slumps and turbidite flows have
Talling, 2006; Muzzi Magalhaes and Tinterri, 2010; Malgesini et al., been reported coming also from active thrust fronts located to the

[BR] sampling site


turbiditic mud-cap and marlstones
Po plain stack of turbidite beds
IT
A
LY

0m
TORTONIAN

Parma 100
0m
Adr

Ap 200
0m
en Bologna
iati

nin
es
cS
ea

Marnoso Arenacea Fm. (MIOCENE)

CR
PRE
GIU Pesaro
Ty

CAB
COR
~ 3000 m
SERRAVALLIAN

CP
r
rh

Firenze PM BR
en
ia

Pisa
n

Verghereto
Se
a

Ape
nni

Isopachs of MA fm. Siena


nes

Approximate location of the [PM]


Apennine front at the Miocene age
Contessa key bed
Outcrop of MA fm. [COR;CAB;GIU;PRE;CP;CR]
LANGHIAN

Perugia
Foreland carbonate source
Apennine source
Alpine source [BR]
50 km
Sampled stratigraphic sections A IO key bed B

Fig. 1. A) Map showing the outcrop area and thickness of the Marnoso Arenacea Formation (after Argnani and Ricci Lucchi, 2001; isopachs in the Po plain subsurface from Dondi et al.,
1982) and location of the 8 sections (small solid squares; geographical locations in Table 1) sampled at various sites through 11 different depositional intervals. Sediment sources were
located mainly in the Alps but also the Apennines including foreland carbonate ramps in the south. B) Schematic stratigraphical log of the Marnoso Arenacea Formation (modified
after Mutti et al., 2002).
F. Felletti et al. / Sedimentary Geology 335 (2016) 197–215 199

west of the basin (Ricci Lucchi, 1975). Scarcity of in situ benthic fossils of the kmax axes either parallel or perpendicular to the current direc-
prevents definite conclusions about paleobathymetry (most macrofos- tion depending on the hydrodynamic boundary conditions. Finally,
sils and macroforaminifera quoted in literature appear reworked). an anisotropic fabric with an overlapping, shingling pattern of mag-
Rare agglutinated foraminifera (Vai, 2001) and abundant bioturbations netic grains is called an imbricated fabric (Fig. 2). Following the def-
(McBridge and Picard, 1991; Monaco and Checconi, 2008; Monaco, inition of Baas et al. (2007), the imbrication angle is the angle at
2008) suggest a bathyal depth on the order of ~1000 m. which a prolate or lobate grain leans away from a horizontal position
Siliciclastic turbidites make up about 90% of the fan lobe and basin- in the bedding plane.
plain deposits, and are characterized by mudstone caps (Bouma's Te di- After deposition, the fabric may be affected by compaction, bioturba-
vision) representing 1/3 to 1/2 of each turbidite bed. Carbonate-rich tion, and disruption by migration of trapped fluids and/or gas, or even
marlstones contribute for the remaining 10% of the total sediment vol- tectonic deformation. As pointed out by Parés et al. (1999), there
ume. Turbidites belong to two distinct classes (Ricci Lucchi and seems to be a general agreement in studies dealing with AMS fabrics
Valmori, 1980): (i) thin turbidites arranged in multilayered stacks, in deformed rocks (e.g., Borradaile and Tarling, 1981, 1984; Parés and
which tend to wedge-out and shale-out gradually down-current, and Dinarès-Turell, 1993; Sagnotti and Speranza, 1993; Mattei et al., 1995)
(ii) single turbidite beds of large volume and spatial extent referred to that the first effect of layer-parallel shortening is to group the kmax
as basin-wide turbidites (or megaturbidites), which tend to maintain axes perpendicular to the shortening direction whereas the kmin axes
a virtually constant thickness (and sand/mud ratio) over distances of tend to remain perpendicular to the bedding plane); with further short-
km to tens of km. These megaturbidites are multi-sourced and consti- ening, the kmin axes progressively rotate into the tectonic shortening di-
tute laterally traceable marker beds, such as the carbonate-rich rection, and the magnetic foliation becomes parallel to the flattening
Contessa megabed or the Colombine megaturbidites comprised of hy- plane and mesoscopic cleavage [examples of this sedimentary-to-
brid arenites, skeletal calcarenites, and marlstones. tectonic fabric progression are provided by Housen et al. (1995),
Averbuch et al. (1992), and Parés et al. (1999)]. Hence, in order to dis-
3. The anisotropy of magnetic susceptibility (AMS) criminate between sedimentary and tectonic fabrics, the relationships
between the AMS and the geometry of sedimentary and tectonic fea-
After the pioneering work of Rees (1965), several authors applied tures must be studied in detail.
the AMS to study paleocurrents in sedimentary rocks, obtaining useful We consider the AMS fabric of the studied samples as induced essen-
results especially in sediments of appropriate grain size, usually clay tially by sedimentary (primary) processes rather than tectonic (second-
and fine sand (Galehouse, 1968; Hamilton and Rees, 1970; Argenton ary) processes essentially because we observed a strong correlation
et al., 1975; Taira and Scholle, 1979; Hiscott and Middleton, 1980; between hydrodynamic regimes, as deduced from sedimentary facies
Ledbetter and Ellwood, 1980; Knode et al., 1990; Schieber and analysis, and associated AMS fabrics, as illustrated below (see also
Ellwood, 1993; Tarling and Hrouda, 1993; Hiscott et al., 1997; Liu Dall'Olio et al., 2010, 2013).
et al., 2001, 2005; Parés et al., 2007; Veloso et al., 2007; Baas et al.,
2007). The AMS reflects the alignment of magnetic grains attained in 4. Sampling and methods
the final stages of sediment transport, with the maximum susceptibility
axis, kmax, and the minimum susceptibility axis, kmin, approximating the Eight stratigraphic sections (Fig. 1A; coordinates in Table 1)
preferred orientation of the longest and shortest magnetic grain axes, (Talling et al., 2012, 2013) have been sampled at different sites com-
respectively (e.g., Hamilton and Rees, 1970; Taira and Scholle, 1979; prising 11 different depositional intervals (sensu Harms, 1975;
Tarling and Hrouda, 1993; Borradaile et al., 1999, Baas et al., 2007). Walker, 1984). These sections straddle altogether a ~ 60 m-thick in-
This method is based on the fact that a current is able to orient paramag- terval of the Marnoso Arenacea Formation, essentially comprised be-
netic grains (e.g., phyllosilicates, olivines, pyroxenes, anphiboles) and tween the Contessa and the Fiumicello marker beds (Carta Geologica
ferromagnetic (sensu lato) grains (e.g., magnetite, goethite, hematite), dell'Appennino Emiliano-Romagnolo 1:10.000) (Fig. 1B). In order to
and that the resulting AMS ellipsoid (Fig. 2) reflects the orientation obtain a direct estimate of flow directions, we measured with a mag-
imparted by the current to such grains (e.g., Ellwood, 1980; Lowrie netic compass the orientation of flute marks at the base of turbidite
and Hirt, 1987; Taira, 1989; Sagnotti and Meloni, 1993; Parés et al., beds. Flute marks, represented by protuberances in the sandstone
2007); naturally, these grains should be proven to be primary (syn-sed- bed or hollows in the underlying mud layer (Rich, 1950), are pro-
imentary) in origin rather than secondary (diagenetic), and in this re- duced by high energy currents; they show long shape axes oriented
spect, thin section analyses (see below) show that there is no parallel to the average current direction (Kuenen, 1957) and asym-
evidence of diagenetic mineral alteration that may affect the magnetic metric shapes that allow to distinguish between up- and down-
fabric of the studied rocks. current directions.
For example, primary phyllosilicates tend to accumulate in still Samples for AMS and additional paleomagnetic analyses were col-
water with their short shape axes perpendicular to the bedding lected with a water-cooled rock drill and oriented with a magnetic com-
plane, and as a result, an oblate mineralogical fabric develops pass. Selected marker beds were sampled in multiple sections to study
(well-developed foliation). The AMS fabric mimics this because in the magnetic fabric up- and down-current within the same flow unit.
phyllosilicates, the short shape axis corresponds to the crystallo- Cylindrical samples were reduced in the laboratory to standard
graphic c-axis as well as to the minimum susceptibility direction, ~10 cm3 specimens. AMS analyses were carried out on 551 specimens
and therefore, a magnetic foliation (defined by the plane containing with a KLY-3 Kappabridge. A susceptibility tensor was then fit to the
kmax and kint) develops parallel to the depositional surface (Ellwood, data by means of least square analysis, and the errors of the fit were cal-
1980; Tarling and Hrouda, 1993; Parés et al., 2007). When currents culated using multivariate statistics (Agico KLY-3 User's Guide, Ver. 2.2
are present (Fig. 2), hydraulic forces control grains alignment (cur- Nov., 1998). Susceptibility tensors were subsequently rotated into tilt-
rent-induced fabric; Shor et al., 1984). The long shape axis of elon- corrected coordinates using site-mean bedding attitudes, and then plot-
gated paramagnetic grains, as well as of large ferromagnetic grains, ted on stereographic projections. We also obtained for each site the
lines up either parallel or perpendicular to the current direction in mean magnetic lineation L = kmax/kint (Balsley and Buddington, 1960)
case of moderate (b 1.2 m/s) or high (N 1.2 m/s) hydrodynamic re- and magnetic foliation F = kint/kmin (Stacey et al., 1960) (Table 2). Oblate
gimes, respectively (Allen, 1984). The AMS fabric reflects this be- (foliated) and prolate (lineated) fabrics were plotted on Flinn-type dia-
cause in such elongated grains, the maximum susceptibility axis grams (Hrouda, 1982; Tarling and Hrouda, 1993). All paleomagnetic
tends to lie broadly along the particle length. Consequently, a sedi- analyses were carried out at the Alpine Laboratory of Paleomagnetism
mentary magnetic lineation can develop as revealed by a clustering of Peveragno, Italy.
200 F. Felletti et al. / Sedimentary Geology 335 (2016) 197–215

Flow-transverse (rolling) fabric


b-axis b-axis
c-axis c-axis

a-axis a-axis
Flow direction Flow direction
Non-imbricated grain fabric Imbricated grain fabric

Nord Flow direction Nord Flow direction

k max
k int
k min

Non-imbricated magnetic fabric Imbricated magnetic grain fabric

Flow-aligned fabric
a-axis a-axis

c-axis c-axis
b-axis b-axis
Flow direction Flow direction

Non-imbricated grain fabric Imbricated grain fabric

Nord Flow direction Nord


Flow direction

k max
k int
k min

Non-imbricated magnetic fabric Imbricated magnetic grain fabric

Horizontal fabric Flow-oblique fabric


c-axis a-axis

b-axis

Flow direction Flow direction


Disperse grain fabric Non-imbricated grain fabric

Nord Flow direction Nord Flow direction

k max
k int
k min

Disperse magnetic fabric Non-imbricated magnetic fabric

Fig. 2. Main types of anisotropic grain shape fabrics. The orientation of elongated grains and the orientation of the three principal orthogonal axes in upper hemisphere stereograms are
drawn. Non-imbricated and imbricated sub-fabrics are shown in the left and right panels, respectively. Horizontal fabric is non-imbricated by definition; a = long grain axis; b = grain axis
of intermediate length; c = short grain axis.
F. Felletti et al. / Sedimentary Geology 335 (2016) 197–215 201

Table 1
Geographic location of sampling sites.

Sampling section Name Latitude Longitude

COR Corniolo lat = 43°54′30.54″N lon = 11°47′41.80″E


PRE Premilcuore lat = 43°58′45.52″N lon = 11°46′53.31″E
CP Castelpriore lat = 43°49′29.47″N lon = 12°7′21.00″E
CAB Cabelli lat = 43°55′50.87″N lon = 11°51′7.40″E
GIU Giumella lat = 43°58′7.75″N lon = 11°45′44.36″E
CR Castel del Rio lat = 44° 9′42.38″N lon = 11°27′51.41″E
BR Bagno di Romagna lat = 43°49′27.92″N lon = 11°57′30.01″E
PM Passo dei Mandrioli lat = 43°48′40.52″N lon = 11°56′0.27″E

5. Turbidite facies and ams fabrics by well clustered kmax axes with an overall mean direction (302°E;
Table 2) slightly departing (by ~ 14°) from the overall mean
To ease comparison between AMS fabrics and paleocurrent direc- paleocurrent direction from flute casts (316°E) (Fig. 6A). This magnetic
tions from flute casts from the various sites, and therefore make infer- fabric is compatible with relatively low-density suspensions with clasts
ences about potential controls of flowing currents on the observed transported by traction and rapidly deposited in flow-aligned (slightly
AMS fabrics, we had to migrate AMS and flute cast data into a common oblique) fabrics. Imbrication angles of ~10° dipping up-current (i.e., to
reference frame, as follows: the NW) have been occasionally observed (sites BO1c, CAB2b, COR4a,
GIU1b; Fig. 7A, B, C, D, respectively). This feature confirms that
1. we averaged flow directions from flute casts (Fig. 3A, B) in tilt- siliciclastic turbidites flowed from the NW to the SE. Carbonatic and hy-
corrected coordinates (i.e., after correction for bedding tilt) for both brid massive sandstones coming from the SE (Fig. 6B), sampled at 3 sites
siliciclastic turbidites and carbonatic-hybrid turbidites, which result- in 3 sections (n = 31 samples; fabric parameters in Table 2), show more
ed to flow, respectively, from 316°E ± 23° (NW) to the SE (Fig. 3C), scattered kmax axes (relative to siliciclastic sandstones) that seem to
and from 147°E ± 19° (SE) to the NW (Fig. 3D); group around a mean direction (127°E) departing by ~ 20° from the
2. then, each sampled site and associated AMS axes have been rotated mean paleocurrent direction from flute casts (147°E) (Fig. 6B).
into these common reference flow directions; for example (Fig. 4), Interpretation: similar disordered/oblique fabrics have been de-
siliciclastic turbidite sites CP2 and COR4a are characterized by actual scribed in several massive sandstones (Baas et al., 2007), and could be
flow directions from flute-casts of 328°E and 293°E, respectively; caused by high viscosity due to high near-bed sediment concentration
hence, actual flow direction and AMS axes from site CP2 have been combined with frequent clast collisions during rapid settling from sus-
rotated in unison by 12° counter-clockwise, while actual flow direc- pension. The different behaviors between siliciclastic and carbonatic
tion and AMS axes from site COR4a by 23° clockwise, into the refer- sandstones are probably related to differences in the minerals that con-
ence mean direction of 316°E. tribute to the magnetic fabrics in the coarse and fine fractions of the sed-
iment and their subsequent post-sedimentary evolution during burial.
5.1. Massive sandstones (Bouma's Ta interval)
5.2. Parallel-laminated sandstones (Bouma's Tb interval)
These sandstones (Fig. 5A) lack stratification and are characterized
by gradual normal grading (Amy and Talling, 2006). They are commonly These sandstones (Fig. 5A, B, C, D) comprise alternations of laminae
interpreted as the product of deposition at rates of suspended load fall- of different grain sizes (mainly fine sands), usually present at the top of
out high enough to suppress tractional transport (Arnott and Hand, stratified planar beds or massive beds with normal grading. Laminae
1989; Allen, 1991; Kneller and Branney, 1995; Shanmugam, 1996). Al- formation is related to traction coupled with fall-out in upper flow re-
though the massive beds used in the present study were selected care- gimes (Lowe, 1982; Ghibaudo, 1992).
fully, it cannot be excluded that some samples were collected from Siliciclastic parallel-laminated sandstones coming from the NW
depositional intervals that appear massive by eye, but were formed in (Fig. 6C), sampled at 14 sites in 7 sections (n = 141 samples; fabric
upper plane-bed or current ripple flow regimes. parameters in Table 2), are characterized by well clustered kmax axes
Siliciclastic massive sandstones coming from the NW (Fig. 6A), sam- with an overall mean direction (301°E) slightly departing (by ~ 15°)
pled at 6 sites in 5 sections (n = 49 samples; Table 2), are characterized from the overall mean paleocurrent direction from flute casts (316°E)

Table 2
AMS parameters for the 11 studied depositional intervals shown in Figs. 5, 6; n: total number of samples analyzed (n); Kmean = mean susceptibility; kmax = maximum suscep-
tibility; kint = intermediate susceptibility; kmin = minimum susceptibility; L = magnetic lineation; F = magnetic foliation; kmax direction: direction of the maximum axes on the
stereoplot; σ = standard deviation.

Depositional interval n Kmean ± σ kmax ± σ kint ± σ kmin ± σ F±σ L±σ


[+10E-6 SI] [+10E-6 SI] [+10E-6 SI] [+10E-6 SI]

1 Siliciclastic massive sandstones 49 154.52 ± 63.01 157.93 ± 64.77 154.62 ± 63.24 151.02 ± 61.08 1.023 ± 0.017 1.022 ± 0.005
2 Carbonatic and hybrid massive sandstones 31 103.39 ± 42.78 104.97 ± 43.95 103.65 ± 42.79 101.55 ± 41.61 1.019 ± 0.008 1.011 ± 0.007
3 Siliciclastic parallel laminated sandstones 141 172.50 ± 43.09 177.21 ± 43.82 173.24 ± 42.98 167.06 ± 42.55 1.038 ± 0.018 1.023 ± 0.007
4 Carbonatic and hybrid parallel laminated 48 75.88 ± 9.15 76.77 ± 9.33 76.03 ± 9.31 74.83 ± 9.15 1.016 ± 0.010 1.010 ± 0.004
sandstones
5 Siliciclastic cross laminated sandstones 40 157.32 ± 31.08 161.53 ± 32.55 158.74 ± 31.98 151.71 ± 28.76 1.044 ± 0.017 1.018 ± 0.006
6 Carbonatic and hybrid cross laminated 34 97.57 ± 19.32 98.79 ± 19.70 97.81 ± 19.42 96.13 ± 18.85 1.017 ± 0.007 1.010 ± 0.003
sandstones
7 Undulated sandstones 32 147.96 ± 56.41 151.70 ± 58.08 148.55 ± 56.54 143.64 ± 54.67 1.033 ± 0.019 1.020 ± 0.008
8 Siliciclastic convoluted sandstones 45 129.87 ± 39.79 132.70 ± 41.62 130.14 ± 40.64 126.17 ± 37.15 1.031 ± 0.019 1.014 ± 0.006
9 Carbonatic and hybrid convoluted 8 128.45 ± 32.43 131.43 ± 34.01 129.67 ± 33.42 124.25 ± 29.87 1.040 ± 0.020 1.013 ± 0.005
sandstones
10 Debrites 30 152.79 ± 13.83 155.81 ± 14.21 152.99 ± 14.06 149.57 ± 13.30 1.023 ± 0.010 1.019 ± 0.009
11 WM beds 93 157.73 ± 26.48 161.37 ± 27.70 158.45 ± 27.07 153.38 ± 24.75 1.032 ± 0.018 1.018 ± 0.005
202 F. Felletti et al. / Sedimentary Geology 335 (2016) 197–215

A B

50 cm

N N
n = 22 n=6
6
31

14
C 7
D

Fig. 3. In situ measurements of flute marks and linear grooves present at the base of turbiditic beds (panels A and B, respectively) were used to reconstruct average tilt-corrected
paleocurrent directions coming from the NW (316°E) in siliciclastic turbidites (panel C) and from the SE (147°E) in carbonatic and hybrid turbidites (panel D); N = geographic north,
n = number of measurements. See text for discussion.

(Fig. 6C); imbrication angles ranging between 10° and 25° and dipping Both siliciclastic (40 samples) and carbonatic cross-laminated (34
up-current (i.e., to the NW) have been observed only at sites CP5, GIU1a, samples) sandstones (fabric parameters in Table 2), are characterized
and PRE2 (Fig. 7E, F, G). Carbonatic and hybrid parallel-laminated sand- by kmax axes that are poorly clustered on the horizontal plane (Fig. 6E,
stones coming from SE (Fig. 6D), sampled at 5 sites (n = 48 samples; F). The overall kmax mean direction (poorly defined) seems to depart
fabric parameters in Table 2), are characterized by well clustered kmax from the overall mean paleocurrent direction from flute casts (by ~39°
axes with an overall mean direction (127°E) slightly departing (by for siliciclastic sandstones coming from the NW and by ~ 5° for the
~ 20°) from the overall mean paleocurrent direction from flute casts carbonatic and hybrid sandstones coming from SE).
(147°E) (Fig. 6D). Interpretation: these poorly clustered AMS ellipsoids are
Interpretation: the substantially flow-aligned magnetic fabric is interpreted as reflecting the particle transport in the form of grain
interpreted to reflect the dominance of rapid sedimentation from flows (i.e. grain avalanche processes; Rees, 1968, 1983; Allen,
suspension on tractional transport of clasts. The aggradation of plane- 1984), related with migrating bed forms. This depositional mecha-
parallel laminations (Rusnak, 1957) implies bed load transport and nism is known to develop flow-aligned fabric and relatively high im-
suspension settling. Bed load grains tend to maintain their major axes brication angles on the leeward slip face of current ripples. Baas et al.
perpendicular to the main flow direction (Baas et al., 2007), whereas (2007) suggested that deposition of suspension load may also con-
particles transported in suspension are typically oriented with their tribute to produce flow-aligned fabric, principally in depositional in-
major axes parallel to the main flow direction . tervals with steeply climbing ripples. Flow-transverse transport
(rolling fabric) can take place on the up-current side of ripples,
where flow velocity increases; however, the lack of flow-transverse
5.3. Cross-laminated sandstones (Bouma's Tc interval) fabric in Bouma's Tc divisions can be explained (Baas et al., 2007)
considering that in most cases this part of the ripple profile is not
These sandstones (Fig. 5C, D, E, F) are characterized by ripple-cross preserved. Flow-aligned fabrics in Bouma's Tc divisions show relative
laminations related to traction coupled with fall-out under lower flow high circular variance (i.e. characterized by kmax axes that are poorly
regimes conducive to ripple formation. They represent the final flow clustered on the horizontal plane) compared to Bouma's Ta and Tb in-
stages, in which flow density and velocity are sufficiently low for tervals, as a result of local variations in flow direction over current
bedform development (Lowe, 1982; Ghibaudo, 1992). ripples with sinuous and linguoid crest lines.
F. Felletti et al. / Sedimentary Geology 335 (2016) 197–215 203

N CP2 N COR4a
8° n = 14 n=6
32

°
293
N


31

k max
k int
k min
paleocurrent
direction
mean
paleocurrent
direction

Fig. 4. Example of procedure followed to rotate tilt-corrected AMS axes into a common reference direction to ease data comparison at the basin scale. Siliciclastic turbidite sites CP2 and
COR4a are characterized by actual flow directions from flute-casts of 328°E and 293°E, respectively, and have been rotated, in unison with their AMS ellipsoids, by 12° counter-clockwise
and 23° clockwise, respectively, into the reference mean flow direction of 316°E. See text for discussion.

Some ripple cross-laminated sandstones sampled in the Castel del explain the coexistence of a broad range of kmax orientations: a flow-
Rio and Passo Mandrioli sections (CR and PM in Fig. 1) show kmax axes transverse orientation (prevailing when the flow becomes strong
aligned at high angles relative to the paleoflow direction measured by enough to lift deposited grains and allow them to roll and jump on
flute casts at the base of the turbidite beds. This discrepancy, confirmed the bedding plane; Schwarzacher, 1963; Johansson, 1964; Hendry,
by measurements of paleoflow direction from ripple crests alignment, 1976) is easily changed into a more stable flow-aligned orientation.
can be related to a partial flow rebound as the result of successive de-
flections of turbidity currents against the basin margins or intrabasinal 5.5. Convoluted sandstones
highs (Muzzi Magalhaes and Tinterri, 2010).
These medium- to fine-grained sandstones (Fig. 5H, I, L) are charac-
5.4. Undulated sandstones terized by the presence of convolute laminae with load structures. In
some cases, thick basal sandstones with undulated and convoluted lam-
These medium-grained sandstones (Fig. 5G) consist of a series of inae pass upward into mudstones through liquefied intervals with
broad, symmetrical undulations, between 20 and 50 cm in wavelength contorted pseudonodules. Convoluted laminae are often associated
and 2 to 4 cm in amplitude. The undulations are ~10 cm-thick and gen- with ripples that indicate paleocurrents moving in the opposite direc-
erally occur above a graded basal division. They grade upward by de- tion of those indicated by flute casts. Muzzi Magalhaes and Tinterri
creasing in amplitude into parallel laminae. The crests of the (2010) interpreted this facies as due to shear stress caused by internal
undulations are roughly parallel to the sole mark directions. They are wave-induced cyclic loading.
interpreted as longitudinal features generated by upper flow regimes These sandstones, sampled at 5 sites from 2 sections (n = 53 sam-
(Walker, 1967). These siliciclastic sandstones, sampled at 4 sites from ples; fabric parameters in Table 2), are characterized by highly dis-
4 sections (n = 32 sample; fabric parameters in Table 2), are character- persed magnetic susceptibility axes (Fig. 6H, I).
ized by most of kmax axes oriented within ~ 15° from the overall mean Interpretation: these results can be explained considering that
paleocurrent direction (Fig. 6G); a subordinate mode occurs at ~90° of through liquefied intervals, clasts tend to be only weakly flow-aligned
the mean paleocurrent direction, consistent with pronounced flow- (e.g., Lindsay, 1968; Hiscott et al., 1997), mainly in the zone of shearing.
transverse fabric. Albeit highly scattered, the susceptibility axes are not randomly distrib-
Interpretation: the gradual upward decrease in amplitude of undula- uted whereby they tend to define a flow-transverse fabric with mean
tions within the beds, and the continuity of lamination across undula- kmax direction rotated by ~ 90° in siliciclastic sandstones and by ~ 51°
tions, indicate a gradually decreasing energy regime during deposition in carbonatic and hybrid sandstones from the reference mean flow di-
(Rees, 1968, 1983; Allen, 1984). This variation in flow velocity can rection from flute casts (Fig. 6H). A similar discrepancy in the same
204 F. Felletti et al. / Sedimentary Geology 335 (2016) 197–215

A B

C D

E F

Fig. 5. Samples for AMS and additional paleomagnetic analyses were collected in different depositional intervals: massive sandstones (A); parallel-laminated sandstones (B,C,D); cross-
laminated sandstones (C,D,E,F); undulated sandstones (G,H); convoluted sandstones (I,L); debrites (M); white marlstone beds (N).

facies has been observed by Muzzi Magalhaes and Tinterri (2010) 5.6. Debrites
between paleocurrent directions determined from flute casts and
from the vergence of convoluted folds, and interpreted as the result These non-graded, matrix-supported sandstones (Fig. 5M) show cha-
of deflection of turbidity currents against the basin margins or otically distributed clasts in a swirly matrix fabric containing a percentage
intrabasinal highs. Imbrications of k max axes by 10°–20° down- of mud varying from very low (b10%) to high (20 to 50%) (Talling et al.,
current observed in several sites (CAB2c, CR4, CR6, CR7; Fig. 7H, I, J, 2004; Amy and Talling, 2006). Their base is usually flat or display shallow
K, respectively) can be related to the presence of shear waves (i.e. cy- linear grooves but lack well-developed flute casts. Correlation of distant
clic wave loading related to shear stress caused by trains of moving sections in the Marnoso Arenacea Formation basin and elsewhere show
internal waves in high-density turbidity currents; Muzzi Magalhaes that both mud-poor (sandy) and mud-rich (muddy) debrites can extend
and Tinterri, 2010) that may affect the imbrication angle of elongat- laterally for up to 60 × 30 km, and pinch-out abruptly in down-flow and
ed grains at or close to the bedding plane, even producing down cur- cross-flow directions in a fashion consistent with en masse deposition
rent imbrications (Sakai et al., 2002). (Talling et al., 2004; Amy and Talling, 2006; Malgesini et al., 2009;
F. Felletti et al. / Sedimentary Geology 335 (2016) 197–215 205

G H

I L

M N

Fig. 5 (continued).

Fonnesu et al., 2015, Southern et al., 2015; Marini et al., 2016). Amy and 5.7. White Marlstone (WM) beds
Talling (2006) interpret debrites as cohesive flows (debris flows) with
sufficiently high sediment concentration (either sand or mud) to inhibit These beds (Fig. 5N) sharply overlie Bouma's Te intervals of
the segregation of particles of different sizes. turbiditic origin from which they are distinguished by their texture,
Debrites have been sampled at 4 sites from 3 sections (n = 30 sam- lighter color, greater carbonate content (25–45%; chiefly from plank-
ples; fabric parameters in Table 2), and show AMS fabrics that are gen- tonic foraminifera and coccoliths, and rare benthic foraminifera), and
erally disordered or sometimes slightly flow-aligned (Fig. 6J). lesser total organic content (~1% T.O.C versus ~2% T.O.C of the Bouma's
Interpretation: these results can be explained considering that in Te divisions). Their thickness is on the order of b20 cm, rarely exceeding
debrites, as in convoluted sandstones (see above), clasts tend to be 50 cm, and the grain size is comprised between mud and silt. They are
only weakly flow-aligned (e.g., Lindsay, 1968; Hiscott et al., 1997), characterized by a massive, speckled, and generally featureless aspect
mainly in the basal zone of shearing (Bouma and Pluenneke, 1975; with rare primary laminations (partially destroyed by bioturbation),
Enos, 1977). Moreover, frequent clast collisions during rapid settling and centimeter- to decimeter-scale color banding reflecting subtle com-
from suspension combined with high flow viscosity contribute to pro- positional variations. The WM beds are widespread in the basin plain
duce a disordered fabric (Baas et al., 2007). and are commonly interpreted as hemipelagites (Mutti and Ricci
206 F. Felletti et al. / Sedimentary Geology 335 (2016) 197–215

SUB-FACIES PHOTOS AMS STEREOPLOTS AMS kmaxROSE DIAGRAMS


N
SILICICLASTIC MASSIVE SANDSTONES
k maximum
n = 49
sites = 6 k intermediate
k minimum
mean paleocurrent
direction

100 mm
A
N
CARBONATIC AND HYBRID MASSIVE SANDSTONES n = 31 k maximum
sites = 3
k intermediate
k minimum
mean paleocurrent
direction

100 mm

B
N
n = 141
SILICICLASTIC PARALLEL-LAMINATED SANDSTONES sites = 14 k maximum
k intermediate
k minimum
mean
paleocurrent direction

10 mm

C
N
CARBONATIC AND HYBRID PARALLEL-LAMINATED n = 48
sites = 5 k maximum
SANDSTONES - paleocurrent from SE
k intermediate
50 mm k minimum
mean paleocurrent
direction

Fig. 6. From left to the right: photographs of sampled turbiditic facies, the associated AMS stereoplots (see Table 2 for data), and rose diagrams of the maximum susceptibility axes (kmax)
directions and associated mean (red lines) for comparison to the reference mean flow directions from flute-casts (blue arrows: 316° ± 23°E for siliciclastic turbidites and from 147° ± 19°E
for carbonatic and hybrid turbidites). The red line corresponds to the arithmetic mean of the Kmax. The axes of magnetic susceptibility (kmax = squares, kint = triangles, kmin = circles) are
plotted in tilt-corrected coordinates. (A) siliciclastic massive sandstones; (B) carbonatic and hybrid massive sandstones; (C) siliciclastic parallel-laminated sandstones; (D) carbonatic and
hybrid parallel-laminated sandstones; (E) siliciclastic ripple cross-laminated sandstones; (F) carbonatic and hybrid ripple cross-laminated sandstones; (G) undulated sandstones;
(H) siliciclastic convoluted sandstones; (I) carbonatic and hybrid convoluted sandstones; (J) debrites; (K) white marlstone beds. N = geographic north; sites = number of sites; n =
number of samples.
F. Felletti et al. / Sedimentary Geology 335 (2016) 197–215 207

SUB-FACIES PHOTOS AMS STEREOPLOTS AMS kmaxROSE DIAGRAMS


N
SILICICLASTIC CROSS-LAMINTED SANDSTONES n = 40
sites = 4 k maximum
k intermediate
k minimum
mean paleocurrent
direction

10 mm

E
N
CARBONATIC AND HYBRID CROSS-LAMINTED SANDSTONES n = 34 k maximum
sites = 4
k intermediate
k minimum
mean paleocurrent
direction

50 mm
F
N
UNDULATED LAMINAE SANDSTONES n = 32
sites = 4 k maximum
k intermediate
100 mm k minimum
mean paleocurrent
direction

G
N
SILICICLASTIC CONVOLUTED SANDSTONES n = 45 k maximum
sites = 4 k intermediate
k minimum
mean paleocurrent
direction

100 mm
H
Fig. 6 (continued).
208 F. Felletti et al. / Sedimentary Geology 335 (2016) 197–215

SUB-FACIES PHOTOS AMS STEREOPLOTS


Fig. 6 (continued). AMS k maxROSE DIAGRAMS
N
CARBONATIC AND HYBRID CONVOLUTED SANDSTONES n=8
sites = 1 k maximum
k intermediate
k minimum
mean paleocurrent
direction

50 mm

I
N
DEBRITES n = 30 k maximum
sites = 4
k intermediate
k minimum
mean paleocurrent
direction

100 mm
J
N
WHITE MARLSTONE BEDS n = 93
sites = 9 k maximum
k intermediate
k minimum

50 mm
K

Fig. 6 (continued).

Lucchi, 1972, 1975; Rupke, 1976; Mutti, 1977, 1979; Mutti and Johns, phyllosilicates, parallel to the mean current direction in the final
1979; Talling et al., 2007; but see Dall'Olio et al., 2010, 2013). stages of transport.
A well-preserved depositional anisotropic fabric with clustered sus-
ceptibility axes is evident in all WM beds (Fig. 6K; fabric parameters in 5.8. Summary of AMS data
Table 2). The kmin axes are consistently vertical or sub-vertical,
i.e., perpendicular or sub-perpendicular to the bedding planes, whereas Tilt-corrected AMS ellipsoids appear to be very well defined in mas-
the kmax axes are consistently horizontal, i.e., parallel to the bedding sive, parallel laminated, and cross-laminated sandstones (Fig. 6;
planes, and oriented NW–SE. In one case (site COR 1b; Fig. 7L), an imbri- Table 2) insofar as mean kmin directions are vertical or sub-vertical
cated fabric with kmax axes plunging to the SE was observed. (i.e., perpendicular or sub-perpendicular to bedding planes) whereas
Interpretation: these results suggest that the WM beds deposited mean kmax directions lie at low angles relative to the reference mean
under weak velocity flows (Dall'Olio et al., 2010, 2013) that oriented paleocurrent directions from flute-casts. In these facies, Flinn-type
the kmax axes of paramagnetic grains, probably slightly elongated plots indicate a well-developed foliation (slightly oblate fabrics) likely
F. Felletti et al. / Sedimentary Geology 335 (2016) 197–215 209

A N n=8
BO1c
B N n=5
CAB2b
C N n=6
COR4a
D N n=9
GIU1b

E N n = 15 F N n = 11 G N n = 15 H N n=8
CAB2c
CP5 GIU1a PRE2

I N n = 11
CR4
J N n = 12
CR6
K N n=7
CR7
L N n=5
COR1b

k max
k int
k min

Fig. 7. Stereographic projections of the principal axes of magnetic susceptibility (kmax = squares, kint = triangles, kmin = circles) in tilt-corrected coordinates of sites characterized by
magnetic imbrication. The blue arrows represent the reference mean paleocurrent directions from flute casts (316° ± 23°E) for siliciclastic turbidites and 147° ± 19°E for carbonatic
and hybrid turbidites). N = geographic north; n = number of samples; BO1c, etc. = site name.

associated to the deposition of phyllosilicates with short shape axes per- depositional processes that disrupted the original current-induced
pendicular to the bedding planes (Fig. 8). Moreover, in these facies a fabric (e.g., dewatering).
small but relatively consistent offset of up to ~15–20° counterclockwise
has been observed between mean kmax directions and mean paleoflow 6. Magnetic mineralogy
directions from flute casts (Fig. 6A, B, C, D, E). Excluding systematic mea-
surement errors, the cause of the observed offset may be sought in small In order to estimate the contribution of ferromagnetic (sensu latu)
differences of flow directions between the erosive, flute cast-generating particles to the AMS, the anisotropy of isothermal remanent magnetiza-
stage of turbidity currents and the subsequent depositional stages dur- tion (AIRM) (Tarling and Hrouda, 1993) was measured on 43 cylindrical
ing which massive, planar, and cross stratifications are formed (assum- samples. Each sample was AF-demagnetized in a field of 50 mT with a
ing substantial flow alignment of kmax axes in these facies). Taira and Molspin tumbler demagnetizer, then magnetized in a direct field of
Scholle (1979), and Clark and Stanbrook (2001) showed similar upward 20 mT with a PUM-1 Agico Pulse Magnetizer, and, finally, the induced
changes in mean orientation of magnetic fabric in Bouma-type turbi- magnetic remanence was measured on a Agico JR6 spinner magnetom-
dites that seems to support our observations. Processes such as local eter; this procedure was applied for each sample along 12 different di-
changes in current direction due to current meandering, or that involve rections. The AIRM data were rotated into tilt-corrected coordinates
currents interaction, lateral confining slopes (Muzzi Magalhaes and using site-mean bedding attitudes, and plotted on stereographic projec-
Tinterri, 2010), or the Coriolis effect on decelerating flows (e.g., Scott, tions for comparison with the AMS diagrams.
1967, Colburn, 1968; Parkash and Middleton, 1970; Yagishita and The AIRM (Fig. 9, left panels; Table 3;) has been studied on a selected
Jopling, 1983), may all have contributed to the observed offset, which set of samples from sites characterized by well-defined AMS ellipsoids.
is presently under investigation. To ease comparison, AMS and AIRM data have been plotted on common
Finally, highly dispersed AMS fabrics dominate disordered facies equal-area steroplots. Sites CP1a (parallel laminated sandstones) and
such as convoluted and undulated sandstones as well as debrites. BO1c (massive sandstones) show clustered maximum, intermediate,
This is in agreement with depositional processes that partially and minimum AIRM axes (red symbols in Fig. 9, left panels) that broadly
prevented grains orientation (e.g., en masse freezing) or early post- coincide with the maximum, intermediate, and minimum susceptibility
210 F. Felletti et al. / Sedimentary Geology 335 (2016) 197–215

1.05000 cryogenic magnetometer. Thermal demagnetization of the NRM re-


vealed the presence in 46 samples of a magnetic component isolated be-
tween room temperature and 250–300 °C and broadly aligned along the
present-day field direction in in situ coordinates (average MAD = 5°).
1.04000
prolate fabric Removal of these initial magnetizations revealed the presence of
scattered component directions with no linear trending to the origin
of the demagnetization axes.
L=K max /K int

1.03000 The apparent absence of primary (i.e., Miocene) magnetic compo-


nent directions prompted us to adopt NRM data from the literature in
order to restore the present-day basin geometry (Fig. 1) to Miocene
3
1 times (Fig. 11) by subtracting post-Miocene Apennine counterclockwise
1.02000 10 7
11 thrust sheet rotations. We corrected average paleoflow directions by
5
applying a clockwise rotation of 29° ± 8° as suggested by Speranza
8 9 et al. (1997) Muttoni et al. (1998, 2000) for the same general area of
2 this study. This correction yielded N–S average paleoflow directions of
1.01000 4
6 Marnoso Arenacea turbidites deposited during the Miocene in a N–S
oriented foredeep associated with the eastward migration of the Apen-
oblate fabric nine thrust fronts (di Biase and Mutti, 2002) (Fig. 11).
1.00000 Within this regional setting, the Marnoso Arenacea basin can be
1.00000 1.01000 1.02000 1.03000 1.04000 1.05000 subdivided into a shallow-marine Alpine Basin in the north, and a
F=K int /K min deep-marine Apennine Basin in the south, separated by the Valle
Salimbene-Bagnolo Ridge (Fig. 11), which is a roughly ESE-trending,
buried feature interpreted as a wrench zone acting as the kinematic
Fig. 8. Flinn-type plot for the 11 studied depositional intervals. See Table 2 for depositional
interval numbering, definition, and magnetic parameters; L = magnetic lineation; F = transfer between the N–S compression in the Alps and the SW–NE com-
magnetic foliation. pression in the Apennines (di Biase and Mutti, 2002). Large volumes of
sand coming from the southern and western Alps in the north (Gandolfi
et al., 1983; Ricci Lucchi, 1986; Roveri et al., 2002) were able to enter the
axes of the AMS ellipsoid (black symbols in Fig. 9), suggesting a clear adjacent and deeper Apennine Basin through turbidity currents that
contribution to the AMS by ferromagnetic (sensu latu) minerals. The crossed the Valle Salimbene-Bagnolo Ridge possibly along fault-
AIRM and AMS of site GIU3c (cross-laminated sandstones) also show controlled depressions. Then, these turbidity currents moved south-
some degree of similarity, whereas in site GIU1b (massive sandstones), ward within the confined foredeep (di Biase and Mutti, 2002), while
the AIRM axes appear scattered relative to the AMS axes. From these at the same time minor volumes of carbonatic and hybrid turbidites
analyses, we conclude that in the investigated facies, there is a variable moved northward upon entering the foredeep from shallow-water car-
contribution of ferromagnetic (sensu latu) grains to the AMS. bonate platforms located along the southern and south-eastern margins
The nature of these grains has been determined by means of isother- of the basin (Gandolfi et al., 1983) (Fig. 11).
mal remanent magnetization (IRM) acquisition curves and thermal de- In addition to southward-flowing siliciclastic turbidites and
magnetization of a 3-component IRM (Lowrie, 1990) (Fig. 9, central and northward-flowing carbonatic and hybrid turbidites, the Marnoso
right panels, respectively). These analyses reveal the presence of a dom- Arenacea basin was also characterized by the presence of N–S contour
inant low coercivity phase with maximum unblocking temperature of currents, which left their testimony in the white marlstone (WM)
~570 °C interpreted as magnetite coexisting with an intermediate coer- beds. These deposits have frequently been interpreted as due to
civity phase with unblocking temperature of ~320 °C interpreted as a hemipelagic settling of fine-grained particles, but according to AMS
sulfide phase. Hematite with high coercivity and maximum unblocking analysis (Dall'Olio et al., 2013; this study) they are interpreted as depos-
temperature of ~ 680 °C is also occasionally present (e.g., sample ited under weak velocity currents. These results prompted Dall'Olio
GIU1.16). et al. (2013) to suggest that the WM beds may have deposited under
In addition, thin sections were cut parallel and perpendicular to the the effects of contour currents and should therefore be referred to as
flow direction (determined by flute casts) for optical scrutiny and image muddy contourites.
analysis using ImageJ software (Rasband, 2008). Image analysis has
been conducted on thin sections in massive sandstones (site CP;
Table 1). Phyllosilicates (muscovite), magmatic lithics, quartz, and feld- 8. Conclusions
spars have been recognized in abundant carbonate matrix (Fig. 10).
Minerals with an anisotropic shape have been selected and divided in The AMS is a very useful fabric analysis technique for quantifying
three different classes of orientation (red objects oriented up-current, flow directions in turbiditic sandstones provided it is performed in
blue object sub-horizontal, green objects oriented down-current; appropriate depositional intervals selected by means of detailed sed-
Fig. 10). imentological analyses and is cross-validated by direct estimates of
In conclusion, according to our analysis paramagnetic phyllosilicates flow directions from sedimentological indicators (before attempting
and ferromagnetic-bearing magmatic lithics, likely containing magne- extrapolations to cases where sedimentological indicators are ab-
tite, dominate the observed current-induced AMS and AIRM fabrics. sent, e.g., in drill cores). Particular advantages of this technique are
the significantly faster measurement time compared to standard
7. Paleogeography of the Marnoso Arenacea basin petrographic fabric analysis and its capability to characterize the ori-
entation of the entire population of (magnetic) grains in three di-
To define amount and sense of vertical-axis tectonic rotations of Ap- mensions, whereas standard image analysis on orthogonal thin
ennine thrust sheets, and restore AMS data to a pre-rotation paleogeog- sections is intrinsically limited to two-dimensional estimates of
raphy, the natural remanent magnetization (NRM) of the sediments grains' orientations.
was studied by thermally demagnetizing a total of 157 fresh specimens In the Marnoso Arenacea Formation, we observed a strong correla-
in increasing steps from room temperature up to 450 °C, and measuring tion between hydrodynamic regimes as deduced from sedimentary
the NRM after each heating step with a 2G Enterprises 755 DC-SQUID facies analysis and associated AMS fabrics, which implies that these
F. Felletti et al. / Sedimentary Geology 335 (2016) 197–215 211

AMS and AIRM stereoplots IRM diagrams ThIRM diagrams


N
120 120
BO1.17

BO1.17
100
MASSIVE SANDSTONE
80

Magnetisation J
Magnetisation J
80
40
60

0 40

20
-40
0
-80 0 100 200 300 400 500 600 700
BO 1c AMS kmax AIRMmax 0 500 1000 1500 2000 2500 T [°C]
n=8 AMS kint AIRMint [mT]
AMS kmin AIRMmin 0.1 T 0.4 T 2.5 T
n=8
N
GIU1.16
6
MASSIVE SANDSTONE

10 GIU1.16
Magnetisation J

Magnetisation J
4
0

-10
0
0 100 200 300 400 500 600 700
GIU 1b AMS kmax AIRMmax 0 500 1000 1500 2000 2500 T [°C]
n=9 AMS kint AIRMint [mT]
n=9 AMS kmin AIRMmin 0.1 T 0.4 T 2.5 T

N
PARALLEL-LAMINATED SANDSTONE

20 14
CP1.5a CP1.5a
12

10
Magnetisation J

Magnetisation J
10
8

6
0
4

2
-10
0
0 100 200 300 400 500 600 700
CP 1a AMS kmax AIRMmax 0 500 1000 1500 2000 2500 T [°C]
n = 11 AMS kint AIRMint [mT]
n = 10 AMS kmin AIRMmin 0.1 T 0.4 T 2.5 T

N
PARALLEL-LAMINATED SANDSTONE

40 CP4.4
CP4.4
20
Magnetisation J

20
Magnetisation J

0 10

-20
0
0 100 200 300 400 500 600 700
CP 4a AMS kmax AIRMmax 0 500 1000 1500 2000 2500 T [°C]
n=9 AMS kint AIRMint [mT]
n=8 AMS kmin AIRMmin 0.1 T 0.4 T 2.5 T

N
CROSS-LAMINATED SANDSTONE

12 GIU3.26
GIU3.26
6
8
Magnetisation J
Magnetisation J

4 4

0
2
-4

-8 0
0 100 200 300 400 500 600 700
GIU 3c AMS kmax AIRMmax 0 500 1000 1500 2000 2500
T [°C]
n=8 AMS kint AIRMint [mT]
n=6 AMS kmin AIRMmin 0.1 T 0.4 T 2.5 T

Fig. 9. AIRM and AMS stereoplots of selected samples (left panels; see Table 3 for data) and associated isothermal remanent magnetization (IRM) acquisition curves (central panels) and
thermal demagnetization of a 3-component IRM (right panels); AMS kmax = black squares, AMS kint = black triangles, AMS kmin = black circles; AIRMmax = red squares; AIRMint = red
triangles; AIRMmin = red circles; N = geographic north; n = number of samples; BO1c, etc. = site name.
212 F. Felletti et al. / Sedimentary Geology 335 (2016) 197–215

Table 3
AIRM and AMS parameters of samples shown in Fig. 9; n: number of samples per site; Kmean = mean susceptibility; kmax = maximum susceptibility; kint = intermediate susceptibility;
kmin = minimum susceptibility; L = magnetic lineation; F = magnetic foliation; σ = standard deviation.

AIRM

Site n Kmean ± σ [+10E-3 SI] kmax ± σ [+10E-3 SI] kint ± σ [+10E-3 SI] kmin ± σ [+10E-3 SI] L±σ F±σ

CP1a 10 29.82 ± 6.54 32.91 ± 7.50 29.88 ± 6.68 26.66 ± 5.51 1.101 ± 0.037 1.119 ± 0.052
CP4a 8 38.35 ± 6.20 42.40 ± 7.78 38.11 ± 6.65 34.54 ± 4.48 1.111 ± 0.056 1.010 ± 0.072
BO1c 8 378.23 ± 103.50 421.46 ± 121.79 380.12 ± 104.70 333.10 ± 85.25 1.108 ± 0.053 1.138 ± 0.049
GIU1b 9 14.82 ± 2.85 16.22 ± 4.98 14.90 ± 3.03 13.34 ± 1.01 1.074 ± 0.085 1.114 ± 0.185
GIU3c 6 9.12 ± 1.77 9.65 ± 1.81 9.19 ± 1.75 8.51 ± 1.76 1.050 ± 0.027 1.084 ± 0.025

AMS

Site n Kmean ± σ [+10E-6 SI] kmax ± σ [+10E-6 SI] kint ± σ [+10E-6 SI] kmin ± σ [+10E-6 SI] L±σ F±σ

GIU3c 8 116.83 ± 5.35 118.10 ± 5.54 116.94 ± 5.31 115.47 ± 5.20 1.010 ± 0.003 1.013 ± 0.004
GIU1b 9 134.80 ± 4.04 137.41 ± 4.08 134.53 ± 4.01 132.46 ± 4.05 1.021 ± 0.002 1.016 ± 0.002
BO1c 8 279.38 ± 48.07 286.34 ± 51.23 280.11 ± 48.23 271.71 ± 44.77 1.021 ± 0.007 1.030 ± 0.008
CP4a 9 76.56 ± 5.59 77.99 ± 5.92 76.92 ± 5.79 74.78 ± 5.07 1.074 ± 0.005 1.028 ± 0.011
CP1a 11 139.15 ± 4.67 143.49 ± 4.83 140.32 ± 4.81 133.66 ± 4.32 1.023 ± 0.001 1.050 ± 0.003

AMS fabrics are primary in origin and the result of sedimentary (rather before attaining a more stable flow-transverse orientation. The diffi-
than tectonic) processes. The following conclusions have been reached: culty to identify a unique transport mode in these depositional inter-
vals makes the reconstruction of paleocurrent directions from AMS
1) Massive, parallel-laminated, and cross-laminated beds contain-
data more challenging, if compared to massive, parallel-laminated
ing paramagnetic phyllosilicates show well-clustered AMS data
and cross-laminated divisions.
with a single dominant fabric type; AIRM data are substantially
3) Highly dispersed AMS fabrics are apparently common in convoluted
similar to AMS data suggesting a contribution to the AMS by fer-
and undulated sandstones as well as in debrites, suggesting deposi-
romagnetic grains (mainly magnetite probably contained in lithic
tional processes that partially prevented grains orientation (e.g., en
grains). A robust correlation between AMS fabric and
masse freezing) or early post-depositional processes that disrupted
paleocurrent directions from sedimentological indicators
the original current-induced fabric (e.g., dewatering).
(e.g., flute casts) has been found in these facies, albeit a small
4) Our AMS analysis coupled with paleomagnetic and geologic data
but relatively consistent offset of up to ~ 15–20° counterclockwise
from the literature helped reconstructing the complex paleogeogra-
has been observed and tentatively explained as due to small var-
phy of the Marnoso Arenacea foredeep prior to rotational deforma-
iations of flow directions between the erosive, flute cast-
tion of the Apennines since the Miocene.
generating stage of turbidity currents and the subsequent deposi-
tion of massive, planar, and cross stratifications.
2) Fabrics oriented at high angle (~40–60°) relative to the paleocurrent
direction from sedimentological indicators (flow-transverse fabric) Acknowledgments
have been observed only in a few samples and can be interpreted
as the result of successive deflections of turbidity currents against G. Malgesini is acknowledged for his help during field work. The re-
the margins of the basin and/or intrabasinal highs, or due to flow ac- viewers and Associate Editor are warmly acknowledged for their helpful
celerations (triggered by, e.g., interactions with floor topography) suggestions. Financial support was provided by FIRB 2009–2012 funds
able to lift grains from the depositional plane and roll them over it to F. Felletti, and IAS grant to E. Dall'Olio.

N
A B C

k max
k int
k min
paleocurrent
direction n = 14

Fig. 10. Thin section of sample CP2. Grains are mostly quartz, muscovite, feldspars, volcanic and carbonatic lithics, and fragments of fossils (planctonic foraminifera). (A) original photo;
(B) photo elaborated throughout Image Analysis techniques. The orientation of the major axis of these grains with respect to the horizontal plane has been measured. Three classes of
orientation have been defined: green and red objects represent grains characterized by an imbrication, and blue objects represent sub-horizontal grains. Muscovite grains appeared
particularly well oriented: in particular, their long axis appear to be sub-parallel to the flow direction and also clearly imbricated.; (C) AMS steroplot of site CP2 (kmax = squares,
kint = triangles, kmin = circles); N = geographic north; n = number of samples; blue arrow = actual flow direction from flute-casts at the base of the sampled bed (328°E).
F. Felletti et al. / Sedimentary Geology 335 (2016) 197–215 213

Structured South-Alpine belt

Parma Tertiary Alpine basin

Valle Salimbene-
Bologna Bagnolo ridge

Pesaro

Ap
Firenze

en

Fle dria f
(A
Verghereto
nin

xu ore
Pisa
ic

10

ra l
00
0
for m

m
20

ramand)
ed

00
Siena

l
ee

p
p

Perugia

CR

PRE
proto-Apennines

COR CAB

CP
PM
BR

IT
AL
Y

~15 Ma Langhian-Serravallian
Isopachs of MA fm. Foreland carbonate source

Approximate location of the Apennine source


Apennine front Alpine source
Outcrop of MA fm.
Sampling site

Fig. 11. Paleogeographic map of the studied area during the middle Miocene showing the main sources and dispersal patterns of sediment within the Marnoso Arenacea foredeep basin.
See text for details and discussion.

References Argenton, H., Bobier, C., Polveche, J., 1975. La mésure de l'anisotropie de susceptibilité
magnetique dans les flysches; application à la recherche des directions des
Agico KLY-3 User's Guide (Ver. 2.2 Nov.), 1998. Modular System for Measuring Magnetic paleocourants. Sedimentary Geology 14, 149–167.
Susceptibility, Anysotropy of Magnetic Susceptibility, and Temperature Variation of Argnani, A., Ricci Lucchi, F., 2001. Tertiary siliciclastic turbidite systems of the Northern
Magnetic Susceptibility. AGICO Advanced Geoscience Instruments CO. Brno, Czech Apennines. In: Vai, G.B., Martini, P. (Eds.), Apennines and Adjacent Mediterranean Ba-
Republic. sins. Kluwer Academic Publishers, pp. 327–350.
Allen, J.R.L., 1984. Sedimentary Structures: their Character and Physical Basis: Develop- Arnott, R.W.C., Hand, B.M., 1989. Bedforms, primary structures and grain fabric in the
ments in Sedimentology, 30 A/B. Elsevier, Amsterdam, pp. 593–663. presence of suspended sediment rain. Journal of Sedimentary Petrology 59,
Allen, J.R.L., 1991. The Bouma division A and the possible duration of turbidity currents. 1062–1069.
Journal of Sedimentary Petrology 61, 2, 291–295. Averbuch, O., Frizon De Lamotte, D., Kissel, C., 1992. Magnetic fabric as a structural
Amy, L.A., Talling, P.J., 2006. Anatomy of turbidites and linked debrites based on long dis- indicator of the deformation path within a fold-thrust structure: a test case
tance, 120 × 30 km) bed correlation, Marnoso Arenacea Formation, Northern Apen- from the Corbieres (NE Pyrenees, France). Journal of Structural Geology 14,
nines, Italy. Sedimentology 53, 161–212. 461–474.
214 F. Felletti et al. / Sedimentary Geology 335 (2016) 197–215

Baas, J.H., Hailwood, E.A., McCaffrey, W.D., Kay, M., Jones, R., 2007. Directional petrolog- Johansson, C.E., 1964. Orientation of pebbles in running water: a laboratory study.
ical characterisation of deep-marine sandstones using grain fabric and perme- Geografiska Annaler 45A, 85–112.
ability anisotropy: methodologies, theory, application and suggestions for Kneller, C., Branney, M.J., 1995. Sustained high-density turbidity currents and the deposi-
integration. Earth-Science Reviews 82, 101–142. tion of thick massive sands. Sedimentology 42, 607–616.
Balsley, J.R., Buddington, A.F., 1960. Magnetic susceptibility anisotropy and fabric of Knode, T.L., Vickers, K.V., Edmiston, C., 1990. Determination and cross verification of
some Adirondack granites and orthogneisses. American Journal of Science 258, paleodip/paleocurrents directions. AAPG Annual Convention with DPA/EMD Divi-
6–20. sions and SEPM, an Associated Society; Technical Program with Abstracts: AAPG Bul-
Boccaletti, M., Calamita, F., Deiana, G., Gelati, R., Massari, F., Moratti, G., Ricci Lucchi, F., letin, v. 74, p. 696 (pp).
1990. Migrating foredeep-thrust belt system in the northern Apennines and southern Kuenen, P.H.H., 1957. Sole markings of graded greywacke beds. Journal of Geology 65,
Alps. Palaeogeography Palaeoclimatology Palaeoecology 77, 3–14. 231–258.
Borradaile, G.J., Tarling, D., 1981. The influence of deformation mechanisms on magnetic Ledbetter, M.T., Ellwood, B.B., 1980. Spatial and temporal changes in bottom-water veloc-
fabrics in weakly deformed rocks. Tectonophysics 77, 151–168. ity and direction from analysis of particle size and alignment in deep-sea sediment.
Borradaile, G.J., Tarling, D., 1984. Strain partitioning and magnetic fabrics in particulate Marine Geology 38, 245–261.
flow. Canadian Journal of Earth Sciences 21, 694–697. Lindsay, J.F., 1968. The development of clast fabric in mudflows. Journal of Sedimentary
Borradaile, G.J., Fralick, P.W., Lagroix, F., 1999. Acquisition of anhysteretic remanence and Research 38, 1242–1253.
tensor subtraction from AMS isolates true palaeocurrent grain alignments. In: Tarling, Liu, B., Saito, Y., Yamazaki, T., Abdeldayem, A., Oda, H., Hori, K., Zaho, Q., 2001.
D.H., Turner, P. (Eds.), Palaeomagnetism and Diagenesis in Sediments 151. Geological Paleocurrent analysis for the Late Pleistocene–Holocene incised-valley fill of the
Society, London, pp. 139–145 (Special Publications). Yangtze delta, China by using anisotropy of magnetic susceptibility data. Marine Ge-
Bouma, A.H., Pluenneke, J.L., 1975. Structural characteristics of debrites from the ology 176, 175–189.
Philippine Sea. Initial Reports of the Deep Sea Drilling Project 31, 497–505. Liu, B., Saito, Y., Yamazaki, T., Abdeldayem, A., Oda, H., Hori, K., Zhao, Q., 2005. Anisotropy
Carruba, S., Casnedi, R., Felletti, F., 2004. From seismic to bed: surface–subsurface correla- of magnetic susceptibility (AMS) characteristics of tide-influenced sediments in the
tions within the turbiditic Cellino Formation (Central Italy). Petroleum Geoscience late Pleistocene–Holocene Changjiang incised valley fill, China. Journal of Coastal Re-
10, 131–140. search 21, 1031–1041.
Clark, J.D., Stanbrook, D.A., 2001. Formation of large-scale shear structures during de- Lowe, D.R., 1982. Sediment gravity flows II: Depositional models with special reference to
position from high-density turbidity currents, Grès d'Annot Formation, south- the deposits of high-density turbidity currents. Journal of Sedimentary Petrology 52,
east France. In: McCaffrey, W.D., Kneller, B.C., Peakall, J. (Eds.), Particulate Gravity 279–298.
Currents 31. International Association of Sedimentologists, Special Publication, Lowrie, W., 1990. Identification of ferromagnetic minerals in a rock by coercivity and
pp. 219–232. unblocking temperature properties. Geophysical Research Letters 17, 159–162.
Colburn, I.P., 1968. Grain fabrics in turbidite sandstone beds and their relationship to sole Lowrie, W., Hirt, A.M., 1987. Anisotropy of magnetic susceptibility in the Scaglia Rossa pe-
mark trends on the same beds. Journal of Sedimentary Petrology 38, 146–158. lagic limestone. Earth and Planetary Science Letters 82, 349–356.
Costa, E., Pialli, G., Plesi, G., 1998. Foreland basins of the Northern Apennines; relation- Maino, M., Decarlis, A., Felletti, F., Seno, S., 2013. Tectono-sedimentary evolution of the
ships with passive subduction of the Adriatic lithosphere. Memorie della Societa Tertiary Piedmont Basin (NW Italy) within the Oligo–Miocene central Mediterranean
Geologica Italiana 52, 595–606. geodynamics. Tectonics 32, 1–27.
Dall'Olio, E., Felletti, F., Muttoni, G., 2010. Hemipelagites, hemiturbidites or muddy Malgesini, G., Talling, P., Felletti, F., 2009. Bed-by-bed correlations as a test for depositional
contourites? The contribute of magnetic fabric analysis to discriminate depositional models of sediment dispersal into the deep ocean: Marnoso Arenacea formation
mechanisms in fine-grained sediments (Marnoso Arenacea Fm., Miocene, northern (Miocene — Northern Apennines). In: Pascucci, V., Andreucci, S. (Eds.), 27th IAS
Italy). In: Hernández-Molina, J., Stow, D., Llave, E., Rebesco, M., Ercilla, G., Van Rooij, Meeting — Sedimentary of Mediterranean island(s) — Abstracts Book. Editrice
D., Mena, A., Vazquez, J.T., Voelker, A. (Eds.), Deep-water Circulation: Processes & Democratica Sarda (586 p).
Product vol. 11. Geo-Temas, pp. 31–32. Malgesini, G., Talling, P.J., Hogg, A.J., Armitage, D., Goater, A., Felletti, F., 2015. Quantitative
Dall'Olio, E., Felletti, F., Muttoni, G., 2013. Constraints on mechanisms of deep-water mud- analysis of submarine-flow deposit shape in the Marnoso-Arenacea Formation: what
stone deposition in the Marnoso Arenacea Formation (Miocene, Italy) through mag- is the signature of hindered settling from dense near-bed layers. Journal of Sedimen-
netic fabric analysis. Journal of Sedimentary Research 83, 170–182. tary Research 85, 170–191.
di Biase, D., Mutti, E., 2002. The “Proto Adriatic Basin”. In: Mutti, E., Ricci Lucchi, F., Roveri, Marini, M., Patacci, M., Felletti, F., Mccaffrey, W.D., 2016. Fill to spill stratigraphic evolution
M. (Eds.), Revisiting Turbidites of the Marnoso Arenacea Formation and their Basin- of a confined turbidite mini-basin succession, and its likely well bore expression: the
Margin Equivalents: Problems with Classic ModelsExcursion Guidebook. Università Castagnola Fm, NW Italy. Marine and Petroleum Geology 94–111.
di Parma and Eni-Agip Division, 64th EAGE Conference and Exhibition, Florence, Mattei, M., Funicello, R., Kissel, C., 1995. Paleomagnetic and structural evidence for Neo-
Italy, pp. I 1–I 4. gene block rotations in the Central Apennines, Italy. Journal of Geophysical Research
Dondi, L., Mostardini, F., Rizzini, A., 1982. Evoluzione sedimentaria e paleogeografica nella 100 (B9), 17863–17883.
Pianura Padana. Translated Title: Sedimentary and paleogeographic evolution of the McBridge, E., Picard, D., 1991. Facies implications of trichichnus and chondrites in turbi-
Po Plain. Guida alla geologia del margine Appenninico-Padano (Translated Title: Geo- dites and hemipelagites, Marnoso Arenacea Formation (Miocene), Northern Apen-
logic guide to the Apennine-Po margin). pp. 47–58. nines, Italy. Palaios 6, 281–290.
Ellwood, B.B., 1980. Induced and remanent magnetic properties of marine sediments as Monaco, P., 2008. Taphonomic features of paleodictyon and other graphoglyptid trace
indicators of depositional processes. Marine Geology 38, 233–244. fossils in Oligo-Miocene thin-bedded turbidites, northern Apennines, Italy. Palaios
Enos, P., 1977. Flow regimes in debris flow. Sedimentology 24, 1, 133–142. 23, 667–682.
Felletti, F., Bersezio, R., 2010. Quantification of the degree of confinement of a turbidite- Monaco, P., Checconi, A., 2008. Stratinomic indications by trace fossils in Eocene to Mio-
filled basin: a statistical approach based on bed thickness distribution. Marine Petro- cene turbidites and hemipelagites of the Northern Apennines (Italy). Studi Trentini
leum Geology 27, 515–532. di Scienze Naturali, Acta Geologica 83, 133–163.
Fonnesu, M., Haughton, P., Felletti, F., McCaffrey, W., 2015. Short length-scale variability of Mulder, T., Alexander, J., 2001. The physical character of subaqueous sedimentary density
hybrid event beds and its applied significance. Marine and Petroleum Geology 67, currents and their deposits. Sedimentology 48, 269–299.
583–603. Mutti, E., 1977. Distinctive thin-bedded turbidite facies and related depositional environ-
Galehouse, J.S., 1968. Anisotropy of magnetic susceptibility as a paleocurrent indicator; a ments in the Eocene Hecho Group (South-Central Pyrenees, Spain). Sedimentology
test of method. Geological Society of America Bulletin 79, 387–390. 24, 107–131.
Gandolfi, G., Paganelli, L., Zuffa, G.G., 1983. Petrology and dispersal directions in the Mutti, E., 1979. Turbidites et cones sous-marins profonds. In: Homewood, P. (Ed.), Sédi-
Marnoso Arenacea Formation (Miocene, northern Apennines). Journal of Sedimenta- mentation Détritique (Fluviatile, Littorale et Marine). Université de Fribourg, Institut
ry Petrology 53, 493–507. de Géologie, pp. 353–419.
Ghibaudo, G., 1992. Subaqueous sediment gravity flow deposits: practical criteria for their Mutti, E., 1992. Turbidite Sandstones. (Agip). Università di Parma 275 p.
field description and classification. Sedimentology 39, 423–454. Mutti, E., Johns, D.R., 1979. The role of sedimentary by-passing in the genesis of basin
Hamilton, N., Rees, A.I., 1970. The use of magnetic fabric in paleocurrent estimation. In: plain and fan fringe turbidites in the Hecho Group System (South-Central Pyrenees).
Runcorn, S.K. (Ed.), Paleogeophysics. London, Academic Press, pp. 445–464. Società Geologica Italiana, Memorie 18, 15–22.
Harms, J.C., 1975. Depositional Environments as Interpreted from Primary Sedimentary Mutti, E., Ricci Lucchi, F., 1972. Le torbiditi dell'Apennino settentrionale: introduzione
Structures and Stratification Sequences. Society of Economic Paleontologists and all'analisi di facies. Società Geologica Italiana, Memorie 11, 161–199.
Mineralogists (161 p). Mutti, E., Ricci Lucchi, F., 1975. Field Trip A-11: Turbidite facies and facies associations. In:
Hendry, H.E., 1976. The orientation of discoidal clasts in resedimented conglomerates, Mutti, E., Parea, G.C., Ricci Lucchi, F., Sagri, M., Zanzucchi, G., Ghibaudo, G., Iaccarino, I.
Cambro-Ordovician, Gaspé, eastern Quebec. Journal of Sedimentary Petrology 46, (Eds.), Examples of Turbidite Facies Associations from Selected Formations of North-
48–55. ern Apennines: International Association of Sedimentologists. 9th International Con-
Hiscott, R.N., Middleton, G.V., 1980. Fabric of coarse deep-water sandstones, Tourelle For- gress, Nice, France, pp. 21–36.
mation, Quebec, Cananda. Journal of Sedimentary Petrology 50, 703–722. Mutti, E., Ricci Lucchi, F., Roveri, M., 2002. Revisiting Turbidites of the Marnoso-arenacea
Hiscott, R.N., Hall, F.R., Pirmez, C., 1997. Turbidity-current overspill from the Amazon Formation and their Basin-Margin Equivalents: Problems with Classic Models. Excur-
channel: texture of the silt/sand load, paleoflow from anisotropy of magnetic suscep- sion Guidebook. Workshop organized by Dipartimento di Scienze della Terra
tibility and implications for flow processes. Proceedings of the Ocean Drilling Pro- (Università di Parma) and Eni-Divisione Agip. 64th EAGE Conference and Exhibition,
gram, Scientific Results 155, 53–78. Florence, Italy, p. 120 May 27–30.
Housen, B., Van Der Pluijm, B.A., Essene, E.J., 1995. Plastic behavior of magnetite and high Mutti, E., Bernoulli, D., Lucchi, F.R., Tinterri, R., 2009. Turbidites and turbidity currents from
strains obtained from magnetic fabrics in the Parry Sound shear zone, Ontario Gren- alpine ‘flysch’ to the exploration of continental margins. Sedimentology 56, 267–318.
ville Province. Journal of Structural Geology 17, 265–278. Muttoni, G., Argnani, A., Kent, D.V., Abrahamsen, N., Cibin, U., 1998. Paleomagnetic evi-
Hrouda, F., 1982. Magnetic anisotropy of rocks and its application in geology and geo- dence for Neogene tectonic rotations in the Northern Appennines, Italy. Earth and
physics. Surveys in Geophysics 5, 37–82. Planetary Science Letters 154, 25–40.
F. Felletti et al. / Sedimentary Geology 335 (2016) 197–215 215

Muttoni, G., Argnani, A., Kent, D.V., Abrahamsen, N., Cibin, U., 2000. Paleomagnetic evi- Schwarzacher, W., 1963. Orientation of crinoids by current action. Journal of Sedimentary
dence for a Neogene two-phase counterclockwise tectonic rotation in the Northern Petrology 33, 580–586.
Apennines (Italy). Tectonophysics 326, 241–253. Schwehr, K., Driscoll, N., Tauxe, L., 2007. Origin of continental margin morphology:
Muzzi Magalhaes, P., Tinterri, R., 2010. Stratigraphy and depositional setting of slurry and submarine-slide or downslope current-controlled bedforms, a rock magnetic ap-
contained (reflected) beds in the Marnoso Arenacea Formation (Langhian- proach. Marine Geology 240, 19–41.
Serravallian) Northern Apennines, Italy. Sedimentology 57, 1685–1720. Scott, K.M., 1967. Intra-bed palaeocurrent variations in a Silurian flysch sequence,
Nilsen, T.H., Shew, R.D., Steffens, G.S., Studlick, J.R.J., 2007. Atlas of deep-water outcrops. Kirkcudbrightshire, Southern Uplands of Scotland. Scottish Journal of Geology 3,
AAPG Studies in Geology 56 (504 p). 268–281.
Novak, B., Housen, B., Kitamura, Y., Kanamatsuc, T., Kawamura, K., 2014. Magnetic fabric Shanmugam, G., 1996. High-density turbidity currents; are they sandy debris flows? Jour-
analyses as a method for determining sediment transport and deposition in deep nal of Sedimentary Research 66, 2–10.
sea sediments. Marine Geology 356, 19–30. Shor, A.N., Kent, D.V., Flood, R.D., 1984. Contourite or turbidite? Magnetic fabric of fine-
Parés, J.M., Dinarès-Turell, J., 1993. Magnetic fabric in two sedimentary rock types from grained Quaternary sediments, Nova Scotia continental rise. In: Stow, D.A.V., Piper,
the Southern Pyrenees. Journal of Geomagnetism and Geoelectricity 45, 193–205. D.J.W. (Eds.), Fine-Grained Sediments; Deep- Water Processes and Facies vol. 15.
Parés, J.M., van der Pluijm, B.A., Dinarès-Turell, J., 1999. Evolution of magnetic fabrics dur- Geological Society of London, Special Publication, pp. 257–273.
ing incipient deformation ofmudrocks (Pyrenees, Northern Spain). Tectonophysics Southern, S.J., Patacci, M., Felletti, F., McCaffrey, W.D., 2015. Influence of flow containment
307, 1–14. and substrate entrainment upon sandy hybrid event beds containing a co-genetic
Parés, J.M., Hassold, N.J.C., Rea, D.K., van der Pluijm, B.A., 2007. Paleocurrent directions mud-clast-rich division. Sedimentary Geology 321, 105–122.
from paleomagnetic reorientation of magnetic fabrics in deep-sea sediments at the Speranza, F., Sagnotti, L., Mattei, M., 1997. Tectonics of the Umbria–Marche–Romagna arc
Antarctic Peninsula Pacific margin (ODP Sites 1095, 1101). Marine Geology 242, (central-northern Apennines, Italy): new paleomagnetic constraints. Journal of Geo-
261–269. physical Research 102, 3153–3166.
Parkash, B., Middleton, G.V., 1970. Downcurrent textural changes in Ordovician turbidite Stacey, F.D., Joplin, G., Lindsay, J., 1960. Magnetic anisotropy and fabric of some foliated
greywackes. Sedimentology 14, 259–293. rocks from SE Australia. Geofisica Pura e Applicata 47, 30–40.
Pickering, K.T., Hiscott, R.N., Kenyon, N.H., Ricci Lucchi, F., Smith, R.D.A., 1995. Atlas of Taira, A., 1989. Magnetic fabric and depositional processes. In: Taira, A., Masuda, F. (Eds.),
Deep Water Environments: Architectural Style in Turbidite Systems. Chapman & Sedimentary Facies in the Active Plate Margin. Terra Scientific Publishing, Tokyo,
Hall, New York, pp. 303–306. pp. 44–77.
Rasband, W.S., 2008. ImageJ. US National Institutes of Health. Bethesda, Maryland, USA Taira, A., Scholle, P.A., 1979. Deposition of resedimented sandstone beds in the Pico For-
(http://rsb.info.nih.gov/ij/index.html). mation, Ventura Basin, California, as interpreted from magnetic fabric measurements.
Rees, A.J., 1965. The use of anisotropy of magnetic susceptibility in the estimation of sed- Bulletin Geological Society of America 90, 952–962.
imentary fabric. Sedimentology 4, 257–283. Talling, P.J., Amy, L.A., Wynn, R.B., Peakall, J., Robinson, M., 2004. Beds comprising debrite
Rees, A.J., 1968. The production of preferred orientation in a concentrated dispersion of sandwiched within co-genetic turbidite: origin and widespread occurrence in distal
elongated and flattened grains. Journal of Geology 76, 457–465. depositional environments. Sedimentology 51, 163–194.
Rees, A.J., 1983. Experiments on the production of transverse grain alignment in a sheared Talling, P.J., Amy, L.A., Wynn, R.B., Blackbourn, G., Gibson, O., 2007. Evolution of turbidity
dispersion. Sedimentology 30, 437–448. currents deduced from extensive thin turbidites: Marnoso Arenacea Formation (Mio-
Ricci Lucchi, F., 1975. Depositional cycles in two turbidite formations of northern Apen- cene), Italian Apennines. Journal of Sedimentary Research 77, 172–196.
nines (Italy). Journal of Sedimentary Petrology 45, 3–43. Talling, P.J., Malgesini, G., Sumner, E.J., Amy, L.A., Felletti, F., Blackbourn, G., Nutt, C.,
Ricci Lucchi, F., 1986. The Oligocene to recent foreland basins of the Northern Apennines. Wilcox, C., Harding, I.C., Khan, S., 2012. Planform geometry, stacking pattern, and
In: Allen, P.A., Homewood, P. (Eds.), Foreland BasinsIAS Spec. Publ. vol. 8. Blackwell extra-basinal origin of low-strength and intermediate-strength cohesive debris flow
Scientific, Oxford, pp. 105–139. deposits in the Marnoso-Arenacea Formation. Geosphere 8 (6), 1–24.
Ricci Lucchi, F., Valmori, E., 1980. Basin-wide turbidites in a Miocene, over-supplied deep- Talling, P.J., Malgesini, G., Felletti, F., 2013. Can liquefied debris flows deposit clean sand
sea plain: a geometrical analysis. Sedimentology 27, 241–270. over large areas of sea floor? Field evidence from the Marnoso-Arenacea Formation,
Rich, J.L., 1950. Flow markings, grooving, and intra-stratal crumpling as a criteria for rec- Italian Apennines. Sedimentology 60, 720–762.
ognition of slope deposits, with illustrations from Silurian rocks of Wales. Bulletin of Talling, P., Allin, J., Armitage, D., Arnott, R., Cartigny, M., Clare, M.A., Felletti, F., Covault, J.,
the American Association of Petroleum Geologists 34, 717–741. Girardclos, S., Hansen, H., Hill, P., Hiscott, R., Hogg, A., Clark, E.J.H., Jobe, Z.R., Malgesini,
Roveri, M., Ricci Lucchi, F., Lucente, C.C., Manzi, V., Mutti, E., 2002. Stratigraphy, facies and G., Mozzato, A., Naruse, H., Parkinson, S., Peel, F.J., Piper, D.J.W., Pope, E., Postma, G.,
basin fill history of the Marnoso Arenacea Formation. In: Mutti, E., Ricci Lucchi, F., Rowley, P., Sguazzini, A., Stevenson, C.J., Sumner, E., Sylvester, Z., Watts, C., Xu, J.,
Roveri, M. (Eds.), Revisiting Turbidites of the Marnoso Arenacea Formation and 2015. Key future directions for research on turbidity currents and their deposits. Jour-
their Basin-Margin Equivalents: Problems with Classic ModelsExcursion Guidebook. nal of Sedimentary Research 85, 153–169.
Università di Parma and Eni-Agip Division, 64th EAGE Conference and Exhibition, Tarling, D.H., Hrouda, F., 1993. The Magnetic Anisotropy of Rocks. Chapman & Hall,
Florence, Italy, pp. III 1–III 15. London.
Rupke, N.A., 1976. Sedimentology of very thick calcarenite-marlstone beds in a flysch suc- Vai, G.B., 2001. GSSP, IUGS and IGC: an endless story toward a common language in the
cession, southwestern Pyrenees. Sedimentology 23, 43–65. earth sciences. Episodes 24, 29–31.
Rusnak, G.A., 1957. The orientation of sand grains under conditions of “unidirectional” Veloso, E.E., Anma, R., Ota, T., Komiya, T., Kagashima, S., Yamazaki, T., 2007. Paleocurrent
fluid flow: 1. Theory and experiment. Journal of Geology 65, 384–409. patterns of the sedimentary sequence of the Taitao ophiolite constrained by anisotro-
Sagnotti, L., Meloni, A., 1993. Pleistocene rotations and strain in southern Italy: the exam- py of magnetic susceptibility and paleomagnetic analyses. Sedimentary Geology 201,
ple of the Sant'Arcangelo Basin. Annali di Geofisica 36, 83–95. 446–460.
Sagnotti, L., Speranza, F., 1993. Magnetic fabric analysis of the Plio–Pleistocene clayey Viana, A.R., 2008. Economic relevance of conturites Contourites: In: Rebesco, M.,
units of the Sant'Arcangelo basin, southern Italy. Physics of the Earth and Planetary Camerlenghi, A. (Eds.), Developments in Sedimentology vol. 60. Elsevier,
Interiors 77, 165–176. Amsterdam, pp. 493–510.
Sakai, T., Yokokawa, M., Kubo, Y., Endo, N., Masuda, F., 2002. Grain fabric of experimental Walker, R.G., 1967. Upper flow regime bed forms in turbidites of the Hatch Formation,
gravity flow deposits. Sedimentary Geology 154, 1–10. Devonian of New York State. Journal of Sedimentary Research 37, 1052–1058.
Schieber, J., Ellwood, B.B., 1993. Determination of basinwide paleocurrent patterns in a Walker, R.G., 1984. Shelf and shallow marine sands. In: Walker, R.G. (Ed.), Facies Models,
shale succesion from anisotropy of magnetic susceptibility (AMS): a case of study second ed. Reprint Series vol. 1. Geoscience, Canada, pp. 141–170.
of the Mid-Proterozoic Newland Formation, Montana. Journal of Sedimentary Petrol- Yagishita, K., Jopling, A.V., 1983. Grain fabric of planar cross-bedding formed by lateral ac-
ogy 63, 878–880. cretion, Caledon outwash, Ontario, Canada. Journal of Geology 91, 599–606.

Potrebbero piacerti anche