Sie sind auf Seite 1von 29

INTEGRATED CYCLOSTRATIGRAPHY—THE LATEMAR (MIDDLE TRIASSIC, ITALY) 183

INTEGRATED CYCLOSTRATIGRAPHY OF A MODEL MESOZOIC


CARBONATE PLATFORM—THE LATEMAR (MIDDLE TRIASSIC, ITALY)

RAINER ZÜHLKE
Institute of Geology and Palaeontology, Ruprecht Karl University, Im Neuenheimer Feld, 234, D-69120 Heidelberg, Germany
e-mail: zuehlke@uni-hd.de, www.rainer-zuehlke.de

ABSTRACT: An integrated cyclostratigraphic approach was applied to the 460-m-thick succession in the Latemar platform interior. The
approach uses new high-resolution cyclostratigraphic data from vertical sections, lateral tracing of physical surfaces over the platform top,
new and existing biostratigraphic data, existing isotopic ages from volcanic ash layers, and new spectral analyses in order to develop a
genetic cyclostratigraphic model. Hierarchical cycles include meter-scale shallowing-upward microcycles and 2–6 bundled thinning-
upward macrocycles. Lateral tracing and correlation of microcycles and macrocycles provides a high-resolution 2D architectural model
of the platform interior. The large majority of microcycles and macrocycles is physically persistent over the platform top with only moderate
changes in thickness and internal facies. The platform top showed simultaneous vertical aggradation controlled by low-amplitude, high-
frequency sea-level changes. Tied-in cyclostratigraphic and biochronostratigraphic data indicate that the 619–701 microcycles in the
platform interior include little more than a single ammonoid biozone (Secedensis Zone), that the total time interval is shorter than 4.10 Ma
(average total time interval = 1.88 My), and that the interpolated microcycle period is shorter than 5.85 ky (average interpolated microcycle
period = 2.68 ky). Microcycles cannot be reconciled with precession forcing but reflect sub-Milankovitch forcing. Spectral analysis is based
exclusively on accommodation cycles, which represent the only direct indication of external control on cyclic deposition. Blackman–Tukey
spectral, multi-taper spectral, and harmonic analyses indicate highly similar and significant frequencies and amplitudes which are largely
stationary over all subsets applied to the cyclic series. Ratios and periods indicative of orbital forcing in the Milankovitch band potentially
exist at (very) high significance points with ∆t = 4.2 ky. In the Latemar cyclic succession, basic microcycles represent sub-Milankovitch
forcing (4.2 ky), thinning-upward, 2–6 bundled macrocycles short- and long-precession forcing (18, 21 ky), and higher-order cycle bundles
short and long obliquity (35, 45 ky) as well as short-eccentricity forcing (95–105 ky). Because of significant latitudinal temperature gradients
and seasonal climate differences, the Triassic period held a significant potential for sub-Milankovitch fluctuations in coupled ocean–
atmosphere circulation. They probably triggered low-amplitude, high-frequency changes in sea level and controlled the deposition of sub-
Milankovitch microcycles. Previous studies of the Latemar carbonate platform favored a model-dependent approach based on smaller
cyclostratigraphic datasets from single sections and spectral analyses. The resulting orbital-forcing models could not be reconciled with
the existing biochronostratigraphic framework for the Triassic and the Anisian to Ladinian stages. They left a widely noted deep
disagreement between biochronostratigraphic and cyclostratigraphic time scales. In contrast, the new forcing model of this study is based
on a complete 2D cyclostratigraphic dataset, considers all biochronostratigraphic constraints, and includes time-calibrated spectral
analyses. The model reconciles biochronostratigraphic and cyclostratigraphic time scales. The Latemar cyclic series includes the oldest
explicit sub-Milankovitch signal and the oldest set of both sub-Milankovitch and Milankovitch signals yet observed in the geologic record.

INTRODUCTION form or ramp; (2) the lateral facies variability within individual
cycles; (3) the lateral variations in the degree of amalgamation or
Concepts in Cyclostratigraphy condensation of cycles; and (4) stratal terminations of cycle
packages. Few cyclostratigraphic models were based on plat-
Studies in ancient cyclic carbonate platforms started with the form- or ramp-to-basin transects and included several vertical
pioneer works of Sander (1936), Schwarzacher (1947), and Fischer sections (e.g., Montañez and Osleger, 1993; Harris et al., 1993;
(1964). Especially since the late 1980s to early 1990s a large Goldhammer et al., 1994; Read et al., 1995; Yang and Kominz,
number of detailed studies (e.g., Grotzinger, 1986; Hardie, 1986; 1999). In these studies, the correlation of cycle packages depends
Schwarzacher and Haas, 1986; Strasser, 1988; Read, 1989; largely to entirely on lithofacies correlation, the recognition of
Goldhammer and Harris, 1989; Bond et al., 1991; Osleger and large-scale cycle stacking patterns, and few disconformities or
Read, 1991; Montañez and Read, 1992; Goldhammer et al., 1993) marker beds. High-resolution tracing of physical surfaces (cycle
focused on specific aspects of cyclostratigraphy in shallow ma- tops) and correlations based on detailed small-scale cycle stack-
rine carbonate-platform settings. Of prime interest has been the ing patterns and numerous marker beds has never been per-
recognition of cycle stacking patterns, their relation to orbital formed across a complete carbonate ramp or platform. Spatial
forcing in the Milankovitch band (Milankovitch, 1941; Berger and information on physical cycle continuity and internal facies vari-
Loutre, 1994), and the resulting implications for geological time. ability is decisive for genetic cyclostratigraphic models. Predomi-
Up to now, there have been very few integrated cyclostratigraphic nantly external (allocyclic) control of rhythmic deposition re-
studies incorporating high-resolution (1) cyclostratigraphic data quires that the large majority of sedimentary cycles physically
including the spatial architecture of cycles; (2) model-indepen- extend over the platform top or upper ramp. Predominantly
dent time constraints from biostratigraphy, chronostratigraphy, internal (autocyclic) control of rhythmic deposition is character-
and/or magnetostratigraphy; (3) geochemical proxies like min- ized by progradation from topographic and/or bathymetric highs
eralogy and stable isotopes; and (4) spectral analysis. and by stratal terminations (downlap, onlap) against preexisting
Cyclostratigraphic models are often based on a single vertical surfaces (Ginsburg, 1971; Pratt et al., 1992; Read et al., 1995).
section or well. They include little or no spatial information on (1) Cyclostratigraphic models for pre-Quaternary platform or
the lateral physical continuity of cycles over the carbonate plat- ramp settings routinely use cycle stacking patterns and/or spec-

Cyclostratigraphy: Approaches and Case Histories


SEPM Special Publication No. 81, Copyright © 2004
SEPM (Society for Sedimentary Geology), ISBN 1-56576-108-1, p. 183–211.
184 RAINER ZÜHLKE

tral analyses for time calibration. With few exceptions (Drummond become a model example of Mesozoic platforms because (1) the
and Wilkinson, 1993; Peper and Cloetingh, 1995; Hüssner et al., platform geometries are completely exposed, and (2) the platform
2000), they regard specific cycle stacking patterns as proof for interior shows an extraordinarily well developed cyclicity. A large
orbital forcing in the Milankovitch bandwidth, which triggered number of cyclostratigraphic, biostratigraphic, and chronostrati-
rhythmic deposition via climatic and associated sea-level changes. graphic data exist for the Latemar platform and have led to the
It is usually assumed that the basic meter-scale shallowing- development of contradicting forcing models. This has triggered
upward cycles represent the shortest orbital-forcing signal, i.e., an intense debate, informally called the “Latemar controversy”.
the precession in the Milankovitch bandwidth. For instance, cycle
stacking patterns of five shallowing-upward meter-scale cycles Geological Setting
arranged in thinning-upward trends (macrocycles) indicate ec-
centricity superimposed on precession forcing (Goldhammer et The middle Triassic Latemar platform is situated in the
al., 1993; Goldhammer et al., 1994). For spectral analyses of cyclic central part of the Southern Alps of Northern Italy, commonly
series, this approach requires calibration of the time series to the referred to as the Dolomites. In the middle Triassic, the Dolo-
signal, which is assumed to be the 20 ky precession signal. A mites formed part of a wide continental shelf in the westernmost
number of studies have compared ancient cyclic series to a (sub-) Tethys. Carbonate ramps and deeper marine straits (100–200 m)
recent model time series for calibration purposes (e.g., Fischer, prevailed in the Anisian (Zühlke, 2000; Masetti and Trombetta,
1986; Hinnov and Goldhammer, 1991; Yang and Kominz, 1999; 1998). In the latest Anisian to early Ladinian, isolated carbonate
Preto et al., 2001). This approach basically represents a Milanko- platforms developed (Rüffer and Zühlke, 1995; Gianolla et al.,
vitch-model-dependent calibration. It requires that (1) orbital 1998). They were separated by irregularly shaped basins with
forcing in the Milankovitch band is present and is the dominating water depths of several hundred meters. Upper Anisian to
trigger for rhythmic deposition; (2) each basic cycle reflects a Lower Ladinian platform successions belong to the Schlern Fm.
single beat of orbital forcing; (2) other, non-Milankovitch controls (Sciliar Fm.), time-equivalent basinal successions to the
on rhythmic deposition were not active, although they are known Buchenstein Fm. (Livinallongo Fm.; Bosellini and Rossi, 1974;
to exist in the Quaternary; and (3) the shortest period docu- Gaetani et al., 1981).
mented in the cyclic platform or ramp succession is the preces- The Latemar represents a small isolated carbonate platform in
sional period of approximately 20 ky. the southwestern part of the Dolomites. The diameter of the
Model-independent time calibrations of cyclic ramp/basin platform top ranges between 2.5 and 3.0 km. Dunn (1992),
successions or basic shallowing-upward cycles have rarely been Goldhammer et al. (1993), Harris (1994), and Egenhoff et al. (1999)
performed for pre-Quaternary settings. This was due to relatively described the large-scale platform architecture. The major part of
large uncertainties of isotopic ages and insufficient biostratigraphic the platform interior is built by shallow subtidal to peritidal and
data. However, recent high-resolution age-dating techniques based supratidal cycles. The platform margin, a tepee belt, was little
on single grains (e.g., Steiger et al., 1993; Mundil et al., 1996b; elevated in terms of topography. A narrow transitional zone
Hilgen et al., 1997; Bowring et al., 1998) have reduced the uncertain- separates the platform margin from a discontinuous reef belt,
ties in isotopic ages. The biochronostratigraphic framework of which is situated on the upper foreslope.
thick cyclic successions is important, because cycles are used as a
tool for improving the internal resolution and accuracy of the EXISTING DATASETS AND MODELS
geological time in the succession and for understanding the vari-
ability of other geological factors (e.g., climate) in time. Figure 2 presents a concise overview on the existing and new
Miall (1997, p. 208) has summarized the requirements for the cyclostratigraphic and biochronostratigraphic constraints for the
establishment of orbital forcing: (1) study of numerous vertical Latemar cyclic series. A first set of cyclostratigraphic studies was
sections; (2) demonstration of the long-distance lateral persis- presented by Hardie et al. (1986), Goldhammer (1987),
tence of cycles, e.g., by the tracing of physical sedimentary Goldhammer et al. (1987), Goldhammer and Harris (1989),
surfaces; (2) demonstration of the persistent regularity of cycle Goldhammer et al. (1990), Hardie et al. (1991), Hinnov and
periods and bundling; (3) spectral mapping; (4) radiometric dating Goldhammer (1991), and Goldhammer et al. (1993). They de-
of bentonites; (5) relation of absolute ages to orbital frequencies. scribed the lithofacies and cycle stacking patterns in the platform-
interior succession. Basic fifth-order cycles document an internal
Objectives shallowing-upward trend from subtidal to peritidal or supratidal
conditions. The platform-interior succession was subdivided
This study follows an integrated approach in order to develop into four units based on the occurrence or absence of tepees and
a well constrained genetic cyclostratigraphic model for an iso- cycle stacking patterns (Fig. 2, column A; Chart 1 (in pocket in
lated carbonate platform in the early Mesozoic. It includes: back of book), column A1; Goldhammer et al., 1993, p. 362–365):
(1) Lower Platform Facies (LPF); (2) Lower Cyclic Facies (LCF); (3)
• a sub-meter-scale architectural and facies model of a carbon- Tepee Facies (TF); and (4) Upper Cyclic Facies (UCF).
ate platform based on laterally persistent physical sedimen- The cyclic succession of the Latemar platform interior com-
tary surfaces; prised 488 measured shallowing-upward cycles in the LCF to
• a high-resolution biostratigraphic and chronostratigraphic lower UCF plus an estimated 110 cycles in the upper UCF. Five
framework utilizing existing and new age constraints; basic, fifth-order shallowing-upward cycles constitute a thinning-
• minimum or maximum datasets for key parameters, e.g., upward trend and form macrocycles, which was interpreted as 100
cycle number and period, in order to quantify the cumulative ky eccentricity forcing superimposed on 20 ky precessional forc-
error intervals in the resulting cyclostratigraphic model; ing. Spectral analyses based on a Milankovitch model-dependent
• power and amplitude spectra calibrated to both thickness and time calibration supported this model. The authors assumed an
time. Anisian age for the LPF and a Ladinian age for the LCF to UCF.
Existing time scales (e.g., Harland et al., 1982; Palmer, 1983; Harland
This integrated approach was applied to the Latemar carbonate et al., 1990) proposed a Ladinian stage duration of up to 7 ± 14 My
platform in the Southern Alps (Fig. 1). Within the last 15 years it has (sic, 238 ± 5 My to 231 ± 13 My; Hinnov and Goldhammer, 1991;
INTEGRATED CYCLOSTRATIGRAPHY—THE LATEMAR (MIDDLE TRIASSIC, ITALY) 185

A B 0 100 Munich
Vienna
[km] 1
London
3 2 N
Berlin 7 Bozen
Paris
C 4
see B Geneva

5 Milano Latemar
Madrid Rome
6
carbonate
platform
(see C)

C to Karer-Pass
Passo di Carezza

Fault
? N
slo
Section pe
Erzlahnspitze pla
193 m tf. t
op
MTF to UCF
to Ober-
eggen ?
Section ?
Valle della Vècia
Section 232 m
Biv. Sieff Latemartürme LCF to UCF Section
470 m Spiz de la Cagnota
op

LTF to UTF
e

316 m
slop

tf. t

LPF to MTF
pla

Section
Valsordaspitze ?
C. dei Fuori to Forno
s lo t o p

229 m
LTF to MTF section name
pe
f.

thickness
at

? stratigr. range
pl

Section
Cavignon (lateral tracing)
Rif. Torre 56 m platform top to
di Pisa LTF to LCF slope boundary

0 [km] 0.5
to Alpe di
Pampeago ? Contour Interval: 25 m

FIG. 1.—Location of the Latemar carbonate platform in the Southern Alps of Northern Italy. A) Overview. B) Major structural units
in the Alps including the position of the Southern Alps and the Latemar carbonate platform. C) Topographic map of the
Latemar Mountain Range, with the locations of vertical sections (see Chart 1, A3-A7). The boundary between platform top and
slope refers to LCF to MTF times. Legend for Part B: 1, eastern Austroalpine unit, Mesozoic sediments; 2, eastern Austroalpine
unit, Paleozoic sediments and basement; 3, Penninic unit (eastern Alps); 4, Southern Alpine unit (Southern Alps); 5, Penninic
unit (western Alps); 6, Helvetic unit; 7, External massifs; black, Alpine intrusives.

their table 1). This was sufficient to accommodate the necessary of Brack et al., 1996), which were difficult to project to the
< 14.0 My, but at least 11–12 My, for the Latemar cyclic series. platform interior.
Brack and Rieber (1993) presented the first biostratigraphic The studies of Brack et al. (1996) and Mundil et al. (1996b)
ages for the Latemar from the uppermost LPF. According to included the first biostratigraphic and chronostratigraphic frame-
Brack et al. (1996) the cyclic series covers the upper Reitzi to upper work for the Latemar cyclic series. It raised serious doubts about
Gredleri or basal Archelaus Zone. Mundil et al. (1996b) and Brack the Hinnov and Goldhammer (1991) and Goldhammer et al.
et al. (1996) presented the first radiometric ages for the Latemar (1993) model. Given a maximum duration of < 4.7 My for the LCF
series (Fig. 2, column C). On the basis of biostratigraphic data, to UCF, basic shallowing-upward microcycles could not repre-
they correlated U–Pb ages measured for volcaniclastic layers in sent 20 ky precessional cycles.
the basinal Buchenstein Fm. at the Seceda (Northern Dolomites) De Zanche et al. (1995) presented biostratigraphic data for
to the Latemar. These ages constrained the maximum duration of four levels (TP9 to L3) within the LCF and TF of Goldhammer et
the LCF to UCF interval to < 4.7 My, probably between 2 to 4 My. al. (1993). They assigned the two cyclostratigraphic units (Fig. 2,
However, the correlation of the younger radiometric ages to the column B) to the Avisianum to Crassus subzones, uppermost
Latemar platform interior was loosely constrained. It was based Hungarites to lowermost Nevadites Zone, in the biostratigraphic
on ammonoid faunas in two platform-slope localities (L3 and L4 scheme of Mietto and Manfrin (1995). According to Preto et al.
186 RAINER ZÜHLKE

A B C
Goldhammer (1987), De Zanche et al. (1995) Brack and Rieber
Rela- Goldhammer and Preto et al. (2002) (1993), Brack et al.
tive Hinnov (1991), Gold- (1996), Mundil et al.
Age includes data from
hammer et al. (1993) Goldhammer et al. (1993), Mietto and Manfrin (1996b)
Bio- includes data from Palmer (1995); complemented by data from includes data from Gold-
Zones (1983), Harland et al. (1982) Hardenbol et al. (1998) hammer et al. (1993)
Sub-
Zones Stage Cyclo- Absolute Stage Zo- Sub- Cyclo- Absolute Stage Cyclo- Absolute
Substage stratigr. Age [Ma] Substage ne zone stratigr. Age [Ma] Substage Zone stratigr. Age [Ma]
228.0
Volcanics

esti-
mated

UCF
Longobardian

Arche-

Archelaus
laus
Longo-

Longobardian
bardi
cum 238.0
+0.4

Longobardian
-0.7

Gredleri
UCF

“Ladinian”
Gredleri
232.5 Gredleri
“Ladinian”

“Ladinian”
( not to scale )

Margari-
tosum

MTF
Recuba-

“Fassanian”
rensis
Curionii

Curionii
238.5
Curionii MTF
“Fassanian”

“Fassanian”

Chie-
sense

LCF 233.1
Nevadites

Serpia-
nensis Sece- LCF
densis
? 233.7
“Anisian”

“Illyrian”

Crassus

TF 241.2
240.0 234.3 +0.8
LCF -0.6
Avisia-
Hungarites

LPF
“Anisian”

“Anisian”

num
“Illyrian”

“Illyrian”

LPF 234.7
Reitzi LPF
Reitzi

Assumed duration Duration LPF to upper TF Duration LCF–UCF


Dura- LCF–UCF (LCF to UCF unknown)
tion
1 stage max. 2–3 subzones 3 biozones

FIG. 2.—Cyclostratigraphic, biostratigraphic, and chronostratigraphic constraints on the Latemar platform interior scaled to relative
time. Columns A–D list datasets of previous studies, column E datasets of this study. Gray-shaded areas indicate the
biostratigraphic range of the cyclic succession (LCF–UTF, this study) or parts of its (previous studies). Ages in boldface type
represent isotopic ages measured from single zircons in volcanic ash layers intercalated in the cyclic series (Mundil et al., 1996b;
Mundil et al., 2003). Ages in normal typeface represent interpolated ages from biochronostratigraphic charts (Palmer, 1983;
Harland et al., 1982; Hardenbol et al., 1998; Columns A–B). The basal row indicates the biostratigraphic duration of (parts of) the
cyclic series. Column E ties all biochronostratigraphic information available to this study to the biostratigraphic schemes of (1)
Mietto and Manfrin (1995), de Zanche et al. (1995) (Column E1); (2) Brack and Rieber (1993), Brack et al. (1996) (Column E2). The
biostratigraphic duration of the Latemar cyclic series (LCF–UTF), approximately a single ammonoid biozone, is identical in the
two schemes.
INTEGRATED CYCLOSTRATIGRAPHY—THE LATEMAR (MIDDLE TRIASSIC, ITALY) 187

D E

Preto et al. (2001) This study Rela-


tive
Age
includes data from
includes data from Brack and Rieber (1993), Brack et al. (1996), de Zanche et al. (1995) Bio-
Egenhoff et al. (1999) Zühlke et al. (2000, 2001, 2003), Mundil et al. (2003) Zones
Sub-
Stage Cyclo- Absolute Available biochronostratigraphic data in the scheme of Available biochronostratigraphic data in the scheme Zones
Substage stratigr. Age [Ma] Mietto and Manfrin (1995), Hardenbol et al. (1998) of Brack and Rieber (1993), Brack et al. (1996)

E1 Biochronostratigraphic framework E2 Biochronostratigraphic framework


Biostratigraphic scheme: Biostratigraphic scheme:
Mietto and Manfrin (1995) Brack and Rieber (1993), Brack et al. (1996)
Ammonoid faunas (AF): Ammonoid faunas (AF):
De Zanche et al. (1995) Brack and Rieber (1993), Zühlke et al. (2000)
Chronostratigraphy: Chronostratigraphy:
Mundil et al. (2003) Mundil et al. (2003)
Cyclostratigraphy: Cyclostratigraphy:
Zühlke et al. (2001, 2003), this study Zühlke et al. (2001, 2003), this study

( not to scale )
Stage Zo- Sub- Cyclo- Absolute Ammonoid Absolute Cyclo- Stage
Substage ne zone stratigr. Age [Ma] Faunas (AF) Age [Ma] stratigr. Zone Substage
UTF

“Fassanian”

“Ladinian”
Curionii

Curionii Curionii
“Lower Ladinian”

UTF
“Ladinian”

AF-8
UCF
“Fassanian”

UTF
“Ladinian”

Chie-
sense
? 241.5
241.5 +0.4 UCF
-0.6
Nevadites

Serpia- UCF +0.4


nensis -0.6
AF-6 Sece-
densis
“Anisian”

241.2 MTF
“Illyrian”

+0.7
Crassus MTF 241.2 AF-4 -0.6
+0.7
-0.6
MTF LCF
Avisia-
LCF 242.6 AF-2 242.6
±0.7 ±0.7
Hungarites
“Anisian”

num
“Illyrian”

“Illyrian”
“Anisian”

LTF AF-1 LTF


LPF LPF
Reitzi
Reitzi

Duration ± UCF Duration LCF–UTF Duration LCF–UTF


(LCF to UTF unknown) Dura-
tion
( no biostrat. data ) ± 1 biozone (max. 4 subzones) ± 1 biozone

FIG. 2 (continued).—

(2002), the Latemar cyclic series additionally covers the Serpianensis Zühlke et al. (2000), Zühlke and Bechstädt (2000), and Zühlke
subzone. et al. (2001, 2003) have presented new cyclostratigraphic data,
Egenhoff et al. (1999) proposed a modified subdivision of the spectral analyses, and new and existing biostratigraphic data in
Latemar cyclic series, which was based on the observation that combination with chronostratigraphic data by Mundil et al.
tepees also existed at the top of the LPF and in the UCF (sensu (2003) for the Latemar platform interior. Mundil et al. (2003)
Goldhammer et al., 1993). They introduced two new units, termed have measured high-resolution U–Pb ages from single zircons
the Lower Tepee Facies (LTF) and the Upper Tepee Facies (UTF). from three volcaniclastic ash layers intercalated in the cyclic
188 RAINER ZÜHLKE

succession. According to the new cyclostratigraphic, biostrati- have been demonstrated in four 9-m-thick intervals (Egenhoff
graphic, and chronostratigraphic data the basic shallowing- et al., 1999). Additional information on cycle continuity is based
upward microcycle in the Latemar represents a sub-Milanko- on an entirely statistical approach (Preto and Hinnov, 2000). (3)
vitch signal. Milankovitch forcing, including precession, obliq- Spectral analyses have been applied to only a small portion of
uity, and short eccentricity, potentially exists in the larger-scale the cyclic succession (UCF; Hinnov and Goldhammer, 1991;
cycle bundlings. Preto et al., 2001). (4) biostratigraphic and chronostratigraphic
Preto and Hinnov (2000) compared an 84-m-thick section at data have not been integrated with high-resolution cyclostrati-
the Latemartürme with the section measured by Goldhammer et graphic data.
al. (1987) 1 km to the west at the Erzlahnspitze (in Italian, Cima
Forcellone). Preto et al. (2001) measured a single 160-m-thick NEW DATASETS AND MODELS
cyclostratigraphic section in the uppermost MTF to basal UTF
(sic) segment of the cyclic series at the Latemartürme (in Italian, Cyclostratigraphy
Torri di Latemar) (see Chart 1, column A2). They rejected existing
chronostratigraphic and biostratigraphic constraints in favor of a Lithofacies.—
Milankovitch-model-dependent approach. The resulting cyclo-
stratigraphic model for the UCF includes short and long preces- Five vertical sections (Fig. 1) with thicknesses of 194–469 m
sional, obliquity, and eccentricity forcing. cover the complete preserved cyclic succession of the Latemar
platform interior. Sections include lithofacies, basic shallowing-
The Latemar Controversy upward cycles (microcycles), and large stacking patterns
(macrocycles). Chart 1 includes all five sections (columns A3–A7)
In summary, six forcing models exist for the Latemar cyclic and presents a cyclostratigraphic transect of 4.3 km length over
series: (1) combined eccentricity and precession forcing for the 1:5 the whole Latemar platform interior. Existing and new
stacking pattern between megacycles and basic shallowing-up- biochronostratigraphic data have been tied into this new high-
ward cycles (e.g., Goldhammer et al., 1993); (2) cycles of identical resolution cyclostratigraphic framework on the microcycle level.
periods reflecting sea-level oscillations of less than 8 ky (Brack et Seven lithofacies types (LFT) occur in the Latemar platform-
al., 1996); (3) aperiodic “cycles” (Brack et al., 1996); (4) cycles are interior succession:
not allocyclic and of Milankovitch frequency (in Brack et al., 1996;
Peterhänsel and Egenhoff, 2000); (5) sub-Milankovitch control of • subtidal to shallow subtidal wackestones to grainstones
the basic shallowing-upward microcycles, precession forcing in (LFT 1);
the thinning upward 2–6 bundled macrocycles and obliquity and • peritidal to intertidal dolomitic caps (LFT 2);
short-eccentricity forcing in the larger-scale cycle bundlings (e.g., • intertidal to supratidal tepees (LFT 3);
Zühlke and Bechstädt, 2000; Zühlke et al., 2001, 2003); (6) preces- • intertidal to supratidal floatstones, rudstones, or bindstones
sion forcing of the basic shallowing-upward cycles and obliquity (LFT 4), often associated with residual sediment;
and eccentricity forcing in the large-scale cycle bundlings (e.g., • supratidal red residual sediment (LFT 5), usually with thick-
Preto et al., 2001). Bechstädt et al. (2003) have given an overview nesses below 5 cm;
on the current status of the Latemar controversy. • sheet cracks (LFT 6);
• volcaniclastics (LFT 7).
Open Questions.—
Lithofacies types LFT 1–6 build meter-scale shallowing up-
The Latemar controversy has serious implications for car- ward successions that constitute the basic microcycle in the
bonate-platform models in the Mesozoic, including the follow- Latemar succession. Two to six microcycles constitute a thin-
ing: (1) Were the basic shallowing-upward microcycles in the ning-upward trend and form macrocycles. Larger-scale stack-
Latemar actually triggered by external controls? Similar well ing patterns of microcycles are well visible in some cyclostrati-
developed shallowing-upward cycles exist in many Triassic graphic units and parts of the platform interior. However, a
carbonate platform interiors, e.g., the Dachstein Fm. (Northern systematic approach to cycle stacking patterns requires spectral
Calcareous Austrian Alps; Satterly, 1996; Enos and Samankassou, analyses.
1998). (2) Do basic shallowing-upward cycles in the Latemar Wackestones to grainstones (LFT 1) constitute the major part
reflect Milankovitch forcing? (3) Do 1:4–1:5 cycle bundlings of the basic microcycles in the Latemar. Dolomitic caps (LFT 2),
between macrocycles and basic shallowing-upward microcycles tepees (LTF 3), floatstones–rudstones–bindstones (LFT 4), and
necessarily indicate eccentricity superimposed on precessional residual sediment (LFT 5) characterize cycle tops. Sheet cracks
forcing, as usually proposed? (4) Do indications of Milanko- (LFT 6) are concentrated at or directly below cycle tops. They
vitch forcing in the Latemar on a larger scale actually exist if the occur predominantly in association with tepees (LFT 3) and
basic shallowing-upward cycles do not represent precessional floatstones–rudstones–bindstones (LFT 4). Subordinately, sheet
forcing? cracks occur with dolomitic caps (LFT 2) and residual sediment
(LFT 5). Volcaniclastics (LFT 7) represent primary deposits from
Current Restrictions.— ash clouds.
Cyclostratigraphic analysis in this study relies exclusively on
The published datasets and cyclostratigraphic models for changes in accommodation. In carbonate settings supratidal litho-
the Latemar cyclic series include several restrictions, as follows: facies represents a clear indicator of a decrease in accommoda-
(1) The total number of cycles is unknown. Cyclostratigraphic tion, because in-situ carbonate production cannot build up above
information covering a major part of the cyclic series is derived sea level. In the Latemar platform interior, the vertical succession
from a single section (Goldhammer, 1987). The section pre- of (shallow) subtidal wackestones and grainstones (LFT 1) over-
sented by Preto et al. (2001) covers only one-third of the cyclic lain by intertidal to supratidal tepees (LFT 3), intertidal to su-
succession. (2) The lateral continuity of cycles in the platform pratidal floatstones–rudstones–bindstones (LFT 4), or supratidal
interior is uncertain. Lateral facies and thickness variations residual sediment (LFT 5) indicates a clear decrease in accommo-
INTEGRATED CYCLOSTRATIGRAPHY—THE LATEMAR (MIDDLE TRIASSIC, ITALY) 189

dation. The vertical succession of (shallow) subtidal wackestones to 4 m occur at the tops of macrocycles. Clear evidence for
and grainstones (LFT 1) overlain by peritidal to intertidal dolo- subaerial exposure is rare or is lacking completely. Cycle
mitic caps (LFT 2) represents a further sufficiently reliable indica- stacking patterns change laterally.
tor of a decrease in accommodation. • Lower Cyclic Facies (LCF), thickness 83–100 m, 75–115 cycles
Lithofacies type LFT 1 includes several subtypes: (minimum to maximum numbers for all sections); average
cycle thickness 0.87–1.16 m. Microcycles mainly include
• fine-grained wackestones with bioclasts and peloids (LFT 1a); shallowing-upward subtidal to intertidal (lithofacies type
• packstones to grainstones with peloids and abundant bio- LFT 2) or subtidal to supratidal (lithofacies types LFTs 4, 5)
clasts, especially dasycladaceans (LFT 1b); successions. Thinning-upward macrocycles are more com-
• intensely bioturbated wackestones to packstones with on- plete in the southwestern part of the Latemar platform and
coids, pisoids, and aggregate grains (LFT 1c). are moderately amalgamated in the northeast. Few small-
scale tepees occur and are usually restricted to the tops of
Subtypes of lithofacies type 1 occur in: macrocycles.
• Middle Tepee Facies (MTF), thickness 119–134 m, 201–268
• vertical successions of LFT 1a to 1b to 1c within the lower to cycles (minimum to maximum numbers for all sections except
upper portion of a microcycle, representing a continuous Valle della Veccia); average cycle thickness 0.49–0.52 m.
shallowing-upward trend in the subtidal environment; Microcycles mainly include shallowing-upward subtidal to
• vertical successions of LFT 1b to 1a at the base of a shallowing- supratidal (lithofacies types LFTs 3, 5) successions. Thinning-
upward microcycle, representing an initial deepening-up- upward macrocycles are condensed and grade laterally to
ward trend in the subtidal environment (see Chart 1, column large tepee antiforms. The abundance of single and stacked
A2; Preto et al. 2001, ranks 3 and 4) tepees is generally high but changes laterally over the plat-
• several vertical successions of LFT 1a to 1b to 1c within the form interior. In general, the amount of tepees increases from
lower to upper portion of a microcycle, representing repeated West to East.
shallowing-upward trends in the subtidal environment. • Upper Cyclic Facies (UCF), thickness 169 m, 224–284 cycles
(minimum to maximum numbers for the Latemartürme and
Within the subtidal part of shallowing-upward microcycles, Erzlahnspitze vertical sections, including projected data),
vertical successions of LFTs 1a–c change laterally in an unsystem- average cycle thickness 0.60–0.75 m. Microcycles mainly in-
atic manner over tens to hundreds of meters. Continuous subtidal clude shallowing-upward subtidal to intertidal/supratidal
shallowing-upward trends change laterally to subtidal deepen- (lithofacies types LFTs 2, 4–6) successions. Thinning-upward
ing-upward or to repeated shallowing-upward trends. Lateral macrocycles are usually complete. In many macrocycles, the
and vertical variations of fine-grained wackestones (LFT 1a), uppermost 1–2 microcycle(s) grade laterally into tepee
packstones and grainstones (LFT 1b), and bioturbated wackestones antiforms. Stacked tepees do not exist.
and packstones (LFT 1c) in the subtidal parts of microcycles • Upper Tepee Facies (UTF), thickness 75 m (cut by the recent
reflect local, unsystematic changes in paleobathymetry, water erosional surface at the summit of the Latemartürme), 102–
energy, or paleoecologic conditions. 121 cycles (minimum to maximum numbers for the
Latemartürme section); average cycle thickness 0.62–0.73 m.
Cyclostratigraphic Units.— Microcycles mainly include shallowing-upward subtidal and
intertidal to supratidal (lithofacies types LFTs 3–6) succes-
Cyclostratigraphic units in the Latemar platform interior sions. Thinning-upward macrocycles are complete to con-
have been redefined on the basis of (1) statistical parameters densed. In almost every macrocycle, the uppermost 2–3
(average cycle thickness); (2) cycle stacking patterns and the microcycle(s) grade laterally into tepee antiforms. Stacked
degree of amalgamation or condensation of cycles; (3) relative tepees, equivalent to at least two macrocycles, exist. The UTF
amounts of lithofacies types LFTs 1–6; and (4) changes in accom- develops gradually from the UCF.
modation. Figure 3 shows four 30-m-thick intervals from the
Latemartürme section which are representative of the cyclostrati- Tepees occur over the complete cyclic Latemar succession
graphic units LCF to UTF. Table 1 lists key parameters of the LCF (LCF–UTF), although in differing abundance. The presence or
to UTF in each measured section and compares them to the values absence of tepees does not constitute an adequate feature by
given by Goldhammer (1987) and Goldhammer et al. (1993). which to subdivide the succession. This is in clear contrast to the
Cyclostratigraphic units in the Latemar comprise (Figs. 3–6): previous studies by Goldhammer (1987), Goldhammer et al.
(1993), and Preto et al. (2001). They were based on single sections
• Lower Platform Facies (LPF), total thickness 235–255 m. The and did not take into account lateral variations in cycle stacking
following values refer to the upper 90–100 m of the LPF, patterns or lithofacies over tens to hundreds of meters or over the
which have been measured in detail: 58–70 cycles, average whole platform top. According to Goldhammer (1987) (see Chart
microcycle thickness 1.30–1.50 m. In contrast to the cyclic 1, Column A1) and Goldhammer et al. (1993), tepees occurred
succession (LCF–UTF), cycles in the LPF are usually subtidal exclusively in the TF (MTF of this study). Preto et al. (2001) did not
to shallow subtidal. Upward thinning of microcycles occurs include any tepees in their rank series of the UCF (Chart 1,
over thicknesses of 2–12 m (average 5.50 m). Microcycles are Column A2).
often amalgamated beyond recognition; macrocycles are
usually well developed. Small-scale tepees, peritidal facies, Cycle Numbers.—
or subaerial exposure are rare and restricted to the tops of
macrocycles. The Latemartürme section covers the complete cyclic series
• Lower Tepee Facies (LTF), thickness 28–30 m, 9–10 microcycles and serves as the new reference section for the Latemar platform
(minimum/maximum numbers); average cycle thickness, interior (Chart 1, Column A5; Fig. 6). Thickness of the LCF to
2.84–3.15 m. Thinning-upward microcycles are moderately UTF reaches 456 m. The minimum to maximum number of
amalgamated. Subtidal to peritidal tepees with heights of up cycles is 619 to 701. Numbers of cycles in the Valsordaspitze,
190 RAINER ZÜHLKE

LCF Lower Cyclic Facies MTF Middle Tepee Facies UCF Upper Cyclic Facies UTF Uppper Tepee Facies

• shallow subtidal to peritidal • shallow subtidal to supratidal • shallow subtidal to peritidal • shallow subtidal to supratidal
microcycles microcycles microcycles microcycles
• complete to amalgamated • condensed thinning-upward • largely complete thinning- • complete to condensed thin-
thinning-upward macrocycles macrocycles upward macrocycles ning-upward macrocycles
• few tepees • stacked tepees • tepees • abundant tepees
• number of microcycles = • number of microcycles = • number of microcycles = • number of microcycles =
90–91 201–221 226–286 102–121
• average thickness of micro- • average thickness of micro- • average thickness of micro- • average thickness of micro-
cycles = 1.02–1.03 m cycles = 0.54–0.59 m cycles = 0.63–0.75 m cycles = 0.62–0.73 m

[m]

30

25

20

15

10

0 1 2 3 4 1 2 3 4 1 2 3 4 1 2 3 4

FIG. 3.—Features of cyclostratigraphic units in the Latemar cyclic succession. The figure depicts four 30-m-thick intervals that are
representative of the LCF to UTF. Sections have not been idealized or schematized. Cycle numbers and average microcycle
thicknesses refer exclusively to the Latemartürme section. Legend: 1, maximum number of macrocycles; 2, maximum number
of microcycles; 3, minimum number or macrocycles; 4, minimum number of microcycles; see Chart 1 for color legend.

Erzlahnspitze, and Spiz de la Cagnota sections, where the recent Physical Continuity of Cycles.—
erosional surface cuts deeper into the cyclic succession, ranges
between 621 and 745 (including projections from the Chart 1 shows the internal architecture of the Latemar plat-
Latemartürme section). The error interval for the total mini- form interior at microcycle and macrocycle resolution. Lagoonal
mum number of cycles is < ± 1% (619–629) and < ± 3% (699–745) cyclic successions in the Valsordaspitze, Erzlahnspitze,
for the total maximum number of cycles. The error interval for Latemartürme, and Valle della Veccia areas are partly separated
minimum and maximum number of cycles is < ± 9% (619–745). by platform slope areas (Reiterjochspitze), topographic passes,
Minimum and maximum numbers of cycles account for uncer- and minor faults associated to postdepositional volcanic dikes
tain cycle tops. As in every cyclostratigraphic study, vertical (early late Ladinian). Several techniques were used to analyze the
cycle definition is not always certain where individual shallow- internal architecture of the Latemar platform:
ing-upward microcycles are condensed or amalgamated. Thin
cycle tops built by dolomitic caps (LFT 2) and by floatstones– • physical tracing of 30–50 macrocycle tops in each area. Traced
rudstones–bindstones (LFT 4) have been obliterated in places horizons were tied both to vertical sections and high-resolu-
by postdepositional pressure solution along physical surfaces tion tele lens images and panorama photographs;
or by lateral variations in diagenetic overprint. Cycle tops • within the framework of physically traced horizons, the re-
characterized by small-scale tepees (LFT 3) are difficult to define maining macrocycle tops were visually traced on high-resolu-
in the inter-tepee areas. tion images;
FIG. 4.—Photographic panorama showing the cyclostratigraphic units (LPF–UTF) in the northern part of the Latemar (Fig. 1). It includes the Latemartürme (Fig. 6; Chart 1, Column A5) and Valle della Veccia
(Chart 1, Column A6) sections. Boundaries between cyclostratigraphic units are indicated by thick unbroken lines. Age-dated volcanic ash layers (Mundil et al., 2003) are indicated by thin broken lines.

FIG. 5.—Photographic panorama showing the cyclostratigraphic units (LPF–UTF) in the northeastern part of the Latemar (Fig. 1). It includes the Erzlahnspitze section (Chart 1, Column A4). Boundaries
between cyclostratigraphic units are indicated by thick unbroken lines. Age-dated volcanic ash-layers (Mundil et al., 2003) are indicated by thin broken lines.
TABLE 1.—Cyclostratigraphic data and calculated cycle periods for the Latemar platform interior. Minimum and maximum intervals (∆t = 2.68-5.85 ky; cf. Figure 7,
Columns E1–E2) in the Latemartürme section represent values used for the time calibration in spectral analyses (Figs. 11–12). The trace of the section Valle delle Veccia
approaches the vertical facies boundary between the cyclic platform-interior succession and the massive platform margin at a very oblique angle. Cycle tops are partly
indistinct. Therefore, cycle numbers in the upper part of this section appear to be relatively low and are not representative of the platform interior.

Section Section Goldhammer 1987


This study
Valsordaspitze Erzlahnspitze Latemartürme Valle della Spiz de la Cima Forcellone Goldhammer et al. 1993
Cima Valsorda Cima Forcellone Torri di Latemar Vecia Cagnota Erzlahnspitze

10, 11, 12
complete cyclic series – LCF to UTF cyclic series – LCF to UCF
1 1 1 1 12
thickness [m] 474.20 456.60 455.80 463.60 460.70 321.78 / 420.00 thickness meas. / estim. [m]
1 1 1 1 12
min.-max. number of macrocycles 148 - 176 144 - 176 134 - 167 136 - 160 135 - 156 97 / 119 number of macrocycles meas. / estim.
1 1 1 1 12
min.-max. number of microcycles 621 - 745 629 - 738 619 - 701 579 - 660 621 - 699 488 / 598 number of microcycles meas. / estim.
1 1 1 1 12
min.-max. ø thickness/microcycle [m] 0.64 - 0.76 0.62 - 0.73 0.65 - 0.74 0.70 - 0.80 0.66 - 0.74 0.66 / 0.70 ø thickness/microcycle [m]
1 1 1 1
aver.-max. time interval (Dt=2.68-5.85 ky) [Ma] 1.66 - 4.36 1.69 - 4.32 1.66 - 4.10 1.55 - 3.86 1.66 - 4.09 – –
1 1 1 1 12
probable time interval (Dt=4.2 ky, Fig. 12B) [Ma] 2.61 - 3.13 2.64 - 3.10 2.60 - 2.95 2.43 - 2.77 2.61 - 2.94 9.76 / 11.96 time interval (Dt=20 ky) [My]
11 11
Upper Tepee Facies – UTF – –
2 2 2 2 11 11
thickness [m] – – 74.60 – – – –
2 2 2 2 11 11
min.-max. number of macrocycles – – 21 - 26 – – – –
2 2 2 2 11 11
min.-max. number of microcycles – – 102 - 121 – – – –
2 2 2 2 11 11
min.-max. ø thickness/microcycle [m] – – 0.62 - 0.73 – – – –
2 2 2 2
aver.-max. time interval (Dt=2.68-5.85 ky) [Ma] – – 0.27 - 0.71 – – – –
2 2 2 2 11 11
probable time interval (Dt=4.2 ky, Fig. 12B) [Ma] – – 0.43 - 0.51 – – – –
10
Upper Cyclic Facies – UCF Upper Cyclic Facies – UCF
2 4 7 2 10
thickness [m] – 169.60 169.40 169.20 – 114.18 / 210.00 thickness meas. / estim. [m]
2 4 7 2 10
min.-max. number of macrocycles – 50 - 63 45 - 58 40 - 54 – 22 / 46 number of macrocycles meas. / estim.
2 4 7 2 10
min.-max. number of microcycles – 224 - 284 226 - 268 191 - 235 – 118 / 230 number of microcycles meas. / estim.
2 4 7 2 10
min.-max. ø thickness/microcycle [m] – 0.60 - 0.76 0.63 - 0.75 0.72 -0.89 – 0.96 / 0.96 ø thickness/microcycle [m]
2 4 7 2
aver.-max. time interval (Dt=2.68-5.85 ky) [Ma] – 0.60 - 1.66 0.61 - 1.57 0.51 - 1.37 – – –
2 4 7 2 10
probable time interval (Dt=4.2 ky, Fig. 12B) [Ma] – 0.94 - 1.19 0.95 - 1.13 0.80 - 0.99 – 2.40 / 4.60 time interval (Dt=20 ky) [My]

Middle Tepee Facies – MTF Tepee Facies – TF


3 5 9
thickness [m] 123.90 119.50 118.90 101.50 133.70 116.72 thickness [m]
3 5 9
min.-max. number of macrocycles 60 - 65 55 - 67 50 - 63 44 - 46 60 - 61 59 number of macrocycles
3 5 9
min.-max. number of microcycles 220 - 232 213 - 242 201 - 221 156 - 160 256 - 268 295 number of microcycles
3 5 9
min.-max. ø thickness/microcycle [m] 0.53 - 0.56 0.49 - 0.56 0.54 - 0.59 0.63 - 0.65 0.50 - 0.52 0.40 ø thickness/microcycle [m]
3 5 9
aver.-max. time interval (Dt=2.68-5.85 ky) [Ma] 0.59 - 1.36 0.57 - 1.42 0.54 - 1.29 0.42 - 0.94 0.69 - 1.57 – –
3 5 9
probable time interval (Dt=4.2 ky, Fig. 12B) [Ma] 0.92 - 0.97 0.89 - 1.02 0.84 - 0.93 0.66 - 0.67 1.08 - 1.13 5.90 time interval (Dt=20 ky) [My]

Lower Cyclic Facies – LCF Lower Cyclic Facies – LCF


6 8
thickness [m] 99.80 – 92.90 92.20 83.20 89.73 thickness [m]
6 8
min.-max. number of macrocycles 19 - 24 – 18 - 20 23 - 25 15 - 16 14 number of macrocycles
6 8
min.-max. number of microcycles 97 - 115 – 90 - 91 100 - 111 72 - 75 73 number of microcycles
6 8
min.-max. ø thickness/microcycle [m] 0.87 - 1.03 – 1.02 - 1.03 0.83 - 0.92 1.11 - 1.16 1.24 ø thickness/microcycle [m]
6 8
aver.-max. time interval (Dt=2.68-5.85 ky) [Ma] 0.26 - 0.67 – 0.24 - 0.53 0.27 - 0.65 0.19 - 0.44 – –
6 8
probable time interval (Dt=4.2 ky, Fig. 12B) [Ma] 0.41 - 0.48 – 0.38 - 0.38 0.42 - 0.47 0.30 - 0.32 1.46 time interval (Dt=20 ky) [My]
INTEGRATED CYCLOSTRATIGRAPHY—THE LATEMAR (MIDDLE TRIASSIC, ITALY)

Legend. 1, including projections from Latemartürme reference section; 2, cyclostratigraphic unit has been eroded; 3, present erosional surface at the summit of
Valsordaspitze is at 121.5 m above base MTF; uppermost MTF = 1.1 m projected from Latemartürme section; 4, present erosional surface at the summit of Erzlahnspitze is at
93.5 m above base UCF; upper UCF = 77.6 m projected from Latemartürme section; 5, base of section is at 100.5 m below top MTF; lower MTF = 19.6 m projected from
Latemartürme section; 6, outcrops not suited for safe recognition of cycles; previous section (Goldhammer, 1987) was measured here (LCF = 89.73 m); 7, limit of the cyclic lagoonal
succession to the massive platform margin is at 74.70 m above base UCF, upper UCF = 94.50 m projected from Latemartürme section; 8, base of section is at 42.70 m below
top LCF; lower LCF = 45.50 m projected from Latemartürme section; 9, sparse outcrops (approx. 30 m); the limit of the cyclic lagoonal succession to the massive platform margin
is at 112.20 m above base MTF; uppermost MTF = 18.40 m projected from Latemartürme section; 10, measured (meas.) a thickness of 114.18 m and 118 cycles (Goldhammer,
1987); estimated a total number of 230 cycles and a total thickness of 210 m (Goldhammer et al., 1993); 11, UTF of this study was not defined and measured (Goldhammer, 1987;
Goldhammer et al., 1993); for estimates, see footnote 10; 12, including estimates on cycle number and thickness of the UCF detailed in footnote 10.
193
194 RAINER ZÜHLKE

FIG. 6.—Reference section at the southwestern slope of the Latemartürme (in Italian, Torri di Latemar) with the cyclic series between
the LCF and UTF. The lowermost 25 m of the LCF, the LTF (Lower Tepee Facies) and LPF (Lower Platform Facies) are below the
base of the photograph. The trace of the section is indicated by the 25-m-spaced thickness ticks (see Chart 1, Column A5). Different
parts of the section have been assembled along physically traced cycle tops. Cyclostratigraphic levels of biostratigraphic and
chronostratigraphic data have been measured and projected into the Latemartürme reference section. Only isotopically dated ash
layers and the stratigraphic position of age-indicative ammonoid faunas are shown. Biostratigraphic ages are according to Brack
and Rieber (1993), Brack et al. (1996), Rieber (personal communication), and Zühlke et al. (2000). Isotopic ages of ash layers LAT-
30 to -32 are according to Mundil et al. (2003).

• where the physical tracing between the four areas was impos- fall. No accommodation space was left over from the previous
sible, the correlation of identical macrocycles was based on microcycle at the start of the new microcycle.
clearly identifiable marker horizons and recognition of de- More than 90% of all macrocycle tops in the Latemar platform
tailed cycle stacking patterns; interior show physical continuity without stratal terminations
• within selected macrocycles, individual shallowing-upward across the whole preserved platform interior. The number of
microcycles were traced laterally. thinning-upward microcycles within macrocycles includes sub-
ordinate variations caused by varying degrees of internal amal-
Physical tracing of top surfaces of accommodation cycles is gamation or condensation. In the framework of minimum and
decisive, because these surfaces represent high-resolution time maximum number of cycles, 80–95% of all microcycles within
lines. Boundaries between subtidal lithofacies types (LFTs 1a–c) laterally continuous macrocycles show physical continuity. Both
neither necessarily indicate changes in accommodation nor rep- microcycles and macrocycles include moderate lateral variations
resent time lines. This study exclusively considers accommoda- in thickness, internal facies, diagenetic overprint, and the nature
tion cycles that built up to sea level at the end of each cycle and of cycle tops (lithofacies types LFTs 2–5). Lateral tracing does not
experienced supratidal conditions during peak relative sea-level provide indications of internal progradation within macrocycles.
INTEGRATED CYCLOSTRATIGRAPHY—THE LATEMAR (MIDDLE TRIASSIC, ITALY) 195

In the MTF, packages of 2–5, but usually 3–4, thinning-upward bution). Individual U/Pb ages, analytical procedures, and sta-
microcycles build macrocycles. In the LCF and UCF–UTF, pack- tistical methods are discussed by Mundil et al. (2003). Brack et
ages of 3–6, but usually 4–5, thinning-upward microcycles build al. (1996) and Mundil et al. (1996b) described a specific succes-
macrocycles. In general a well developed layer-cake geometry sion of volcaniclastic layers (Tc–Te) in the basinal Seceda sec-
characterizes the Latemar platform interior. This depositional tion, situated 28 km to the northeast in the northwestern Dolo-
architecture, the lack of stratal terminations (onlap, downlap), mites (Gröden Valley, Val Gardena). Horizon Tc, a crystal tuff,
and progradational features indicate an external (allocyclic) con- represents a regional marker horizon in the lower Secedensis
trol on platform development. Zone and can be traced as far as 130 km to the southwest into the
Lombardian Alps. In the Latemar, the Valsordaspitze section
Biostratigraphy includes a crystal tuff layer in the upper LCF, the biostrati-
graphic age of which is lower Secedensis Zone. Single zircons
Ammonoid faunas in the Latemar platform interior constrain from the Tc horizon in the Seceda section yielded an age of 241.2
the biostratigraphic correlation of the cyclic succession. They occur + 0.8/-0.6 Ma. However, it is still uncertain whether the crystal
on at least eight levels within the cyclic succession. As part of this tuff in the Latemar platform is laterally correlative to the Tc
study, ammonoid faunas described from various platform locali- crystal tuff at the Seceda section.
ties (Brack and Rieber, 1993; De Zanche et al., 1995; Brack et al.,
1996) were projected along traced and correlated macrocycle tops Constraints on the Basic Microcycle
into the Latemartürme reference section (Fig. 6). Two ammonoid and the Latemar Cyclic Succession
faunas bracket the complete cyclic succession. The ammonoid
fauna AF-1 of Brack and Rieber (1993) indicates an upper Reitzi High-resolution cyclostratigraphic, biostratigraphic, and
Zone age for the base of the cyclic succession. The rich ammonoid chronostratigraphic data constrain the average microcycle
fauna AF-8 has recently been found approximately 1 m below the duration in the Latemar cyclic succession, provided that (1)
preserved top of the UTF (Rieber, personal communication; Zühlke isotopic ages reflect true ages of deposition; (2) the individual
et al., 2000). It includes Chieseiceras sp., which indicates uppermost microcycle period remained constant during the LCF to UCF;
Secedensis to (?) lowermost Curionii Zone (cf. Vörös, 1998). (3) each external pulse triggering cyclic deposition is actually
Thus biostratigraphic constraints indicate that the Latemar preserved in the cyclic succession; and (4) each shallowing-
cyclic series (LCF to UTF) is of considerably shorter duration than upward microcycle represents a single cycle of accommoda-
previously assumed. It covers little more than a single ammonoid tion change. The time interval enclosed by the two ash layers
biozone of the late “Anisian” to earliest “Ladinian” instead of the LAT-31 and LAT-32 ranges between 0.9 and 1.1 Ma on average
largest part of the “Ladinian” (Goldhammer et al., 1993, equal to and between 2.3 and 2.4 My at the most. Given the cycle
at least five ammonoid biozones) or three to four biozones (Brack number of 381–410 between the two layers, the average and
et al., 1996). maximum periods of each microcycle are ∆tav = 2.20–2.89 ky
The new biostratigraphic data for the Latemar cyclic series and ∆tmax ≤ 5.61–6.30 ky. Further cyclostratigraphic and spec-
imply that a considerable part of the northeastern slope of the tral analyses in this study are based on cycle periods of ∆tav =
Latemar platform postdates the preserved cyclic platform interior 2.68 ky to ∆tmax ≤ 5.85 ky. Even when maximum uncertainties
(LCF–UTF), which terminates in the uppermost Secedensis to (?) on the ages are applied, microcycle periods cannot be recon-
lowermost Curionii Zone. The ammonoid and Daonella faunas ciled with precession forcing in the Triassic, which requires a
described by Brack et al. (1996; L3, L4) indicated upper Gredleri to cycle period of at least 17.8 ky. Therefore basic-shallowing
Archelaus Zone for the upper part of the northeastern slope succes- upward microcycles (accommodation cycles) represent sub-
sion east of Col Cornon. No platform-interior successions equiva- Milankovitch forcing.
lent to this part of the Latemar slope have been preserved because Two approaches can be applied to extrapolate the total time
of Miocene to recent uplift and erosion of the Southern Alps. interval for the complete cyclic series (base LCF to top UTF).
The first approach applies a linear regression of age as a
Radiometric Stratigraphy function of stratigraphic height extrapolated to the base and
the top of the succession (Mundil et al., 2003). It indicates a
Volcaniclastic ash layers intercalated in the cyclic succession time interval of ≤ 6.3 My if maximum error bounds are applied,
are present at at least eight stratigraphic levels. Lateral tracing of and a mean of 2.2 My. The regression is based on a depositional
ash layers shows that three of them are platform-wide marker rate that has not been corrected for compaction and is aver-
horizons. Four layers pinch out laterally, owing to strong dilution aged over the whole cyclic succession (cf. Zühlke et al., 2003,
or loss by storms. Mundil et al. (2003) have sampled volcanic ash their fig. 4). The second approach (this study) applies an
layers at three cyclostratigraphic levels and localities (Chart 1, extrapolation that is based on the cyclostratigraphic and chrono-
Columns A3, A5): (1) Sample LAT-31 in the lower LCF in the stratigraphic framework. The average and maximum cycle
Valsordaspitze section; (2) Sample LAT-30 in the middle MTF in periods calculated for the succession between LAT-31 and
the Latemartürme section; (3) Sample LAT-32 in the middle to LAT-32 are attributed to all cycles in the cyclic succession. As
upper UCF in the Erzlahnspitze section. On the basis of the a result, the total time interval amounts to ≤ 4.1 My at the most
dataset of this study, the sampling levels were tied to the cyclo- (Fig. 7, Column E2), and 1.88 My (Fig. 7, Column E1) on
stratigraphic framework of the platform. The two sampling levels average. The variant total durations of the cyclic series from
LAT-30 and LAT-31 bracket a minimum of 381 and a maximum the two extrapolations do not affect the genetic interpretation
of 410 shallowing-upward microcycles. of the basic accommodation cycle and superimposed cycle
Mundil et al. (2003) present 206Pb/238U ages based on single- bundles in the Latemar carbonate platform.
zircon techniques with precisions at the 2-3 permil level: (1)
LAT-31; 242.6 ± 0.7 Ma; (2) LAT-30; 241.2 +0.7/-0.6 Ma; (3) LAT- Spectral Analysis
32; 241.5 +0.4/-0.6 Ma (median age based on a normal distribu-
tion of U concentrations and/or Th/U ratios), 241.7 + 1.5/-0.7 Five methods of spectral analysis were applied to the com-
Ma (including additional ages deviating from a normal distri- plete 458-m-thick cyclic series (LCF–UTF) in the Latemartürme
196 RAINER ZÜHLKE

Column A A B C D
Constraints.—Cyclostratigraphy: Goldhammer (1987), De Zanche et al. (1995) Brack and Rieber
Goldhammer and (1993), Brack et al. Preto et al. (2001)
Preto et al. (2002)
single, 322-m-thick section. Bio- Hinnov (1991), Gold- (1996), Mundil et al.
includes data from
stratigraphy: none. Chrono- Time hammer et al. (1993) Goldhammer et al. (1993), Mietto and (1996b)
stratigraphy: none. [Ma] includes data from Palmer Manfrin (1995); complemented by includes data from Gold- includes data from
(1983), Harland et al. (1982) data from Hardenbol et al. (1998) hammer et al. (1993) Egenhoff et al. (1999)
Remarks.—Chronostratigraphic Cyclo- Absolute Cyclo- Sub- Absolute Cyclo- Absolute Cyclostrati- Absolute
Stage Zone Zones
framework was inferred from stratigr. Age [Ma] stratigr. zone Age [Ma] stratigr. Age [Ma] graphic units Age [Ma]
Milankovitch model. Spectral 227.0
analysis was restricted to UCF.
228.0 228.0
Column B Vol-
Constraints.—Cyclostratigraphy: canics

approximate boundaries of 229.0

units according to Goldham- Arche-


esti- Longob. laus
mer (1987) and Goldhammer et 230.0 mated
230.3
al. (1993). Biostratigraphy: am- Gredleri
Gred-
monoid faunas in the LCF–TF 231.0 Margarit.
leri
units, biostratigraphic scheme UCF 231.4
Recubar.
of Mietto and Manfrin (1995). 232.0 Cur-
ionii
Curionii
Chronostratigraphy: none. 232.5 232.6
Remarks.—Absolute ages were not 233.0 Chiesen.
233.1
included in De Zanche et al. “Ladin-
ian” Serpian.
Neva-
dites
(1995). Ages have been added 234.0
?
Crassus
233.7

for comparison because the TF 234.3


LCF Avisian. Hunga-
Mietto and Manfrin (1995) bio- rites
234.7
235.0 Reitzi
stratigraphic scheme has been LPF
TF
incorporated in the time scales
236.0
of Gradstein et al. (1994, 1995)
and Hardenbol et al. (1998).
237.0

Column C Arche-
laus 238.0
Constraints.—Cyclostratigraphy: 238.0 +0.4
? Gred- -0.7
boundaries of units according 238.5
leri 21.7 ky 17.6 ky
to Goldhammer (1987) and 239.0
UCF
LCF
Goldhammer et al. (1993). Bio-
stratigraphy: ammonoid and 240.0 240.0 TF Cur-
ionii
Daonella faunas from the up- LPF “Anisian”

per LPF and the platform slope; 241.0 LCF Sece- 241.2
densis +0.8
biostratigraphic scheme of -0.6 UCF UCF ?
Brack and Rieber (1993). Chro- LPF Reitzi
242.0
nostratigraphy: U–Pb ages
from the northern Dolomites
243.0
(Seceda Section, Buchenstein MTF
Fm.) were correlated to the MTF ?
Latemar on the basis of bio- 244.0
?
stratigraphic data.
Remarks.—The study includes the
Assumed duration Interpolated duration LCF to upper TF Interpolated duration Infered duration ± UCF
first biostratigraphic and Dura- LCF–UCF (LCF to UCF unknown) LCF–UCF (LCF to UTF unknown)
chronostratigraphic framework tion
11–12 Ma, ≤ 14 Ma < 1.6 Ma ≤ 4.7 My, 2–4 Ma 3.20–3.95 Ma
for the Latemar cyclic series.

Column D
Constraints.—Cyclostratigraphy: FIG. 7.—Cyclostratigraphic, biostratigraphic, and chronostratigraphic constraints on the
single 160 m section in ± UCF; Latemar platform interior scaled to absolute time (in full scale). Columns A–D list
cyclostratigraphic scheme ac- constraints of previous studies, and column E, datasets of this study. Gray-shaded
cording to Egenhoff et al. (1999). areas indicate the chronostratigraphic range of the cyclic succession or parts of its.
Biostratigraphy: none. Chrono- Ages in boldface type represent isotopic ages measured from single zircons in
stratigraphy: none. volcanic ash layers intercalated in the cyclic series (Mundil et al., 1996b, 2003). Ages
Remarks.—Existing chronostrati- in normal typeface represent interpolated ages from biochronostratigraphic charts
graphic and biostratigraphic (Palmer, 1983; Harland et al., 1982, 1990; Hardenbol et al., 1998, Columns A–B) or
constraints were rejected in fa- extrapolated ages based on cyclostratigraphic and biochronostratigraphic constraints
vor of a Milankovitch-model- (this study, Column E). The basal row indicates the extrapolated absolute duration of
dependent calibration of the (parts of) the cyclic series (LCF–UTF).

(continued at the top of page 197)


INTEGRATED CYCLOSTRATIGRAPHY—THE LATEMAR (MIDDLE TRIASSIC, ITALY) 197

E time series. Spectral analysis


This study was applied to the uppermost
MTF to the upper UCF, based
includes data from Time on a rank series of four litho-
Brack and Rieber (1993), Brack et al. (1996), [Ma] facies types reflecting paleo-
Zühlke et al. (2000, 2001, 2003), Mundil et al. (2003)
Cyclo-, bio-, and chronostratigraphic Cyclo-, bio-, and chronostratigraphic Cyclo-, bio-, and chronostratigraphic
bathymetric changes. Preto et
constraints with average values constraints with maximum error intervals constraints combined with spectral analysis al. (2001) did not present any
227.0
microcycle numbers. Their
rank series includes at least
182 cycles (cf. Chart 1). On the
228.0
basis of their model of preces-
sion forcing (∆t = 17.7–21.6
229.0
ky) the section has been cen-
tered to the U–Pb age of 241.5
230.0 ± 0.4 My (see Preto et al., 2001,
fig. 1) cited from Zühlke (per-
231.0 sonal communication, 2001).

232.0 Column E
Constraints.—Cyclostratigraphy:
E1 Constraints - average values E2 Constraints - maximum error intervals Constraints and spectral analysis
233.0 complete cyclic Latemar series
E3
(LCF–UTF) measured in five
maximum number of cycles maximum number of cycles maximum number of cycles
average U–Pb ages U–Pb ages with maximum U–Pb ages shifted within their 234.0
sections, 190–470 m thick,
biostratigraphy error intervals respective error intervals across the platform interior.
Ø ∆t microcycle = 2.68 ky biostratigraphy biostratigraphy Biostratigraphy: ammonoid
max. ∆t microcycle ≤ 5.85 ky spectral analysis results (Fig. 12B) 235.0
∆t microcycle = 4.2 ky
faunas from the top, internal
succession and base of the cy-
236.0
Cyclo- Inter- U–Pb Cyclo- Inter- U–Pb Cyclo- Inter- U–Pb clic series described by Brack
stratigr. polated ages Zones stratigr. polated ages Zones stratigr. polated ages Zones
Units ages [Ma] [Ma] Units ages [My] [Ma] Units ages [Ma] [Ma] and Rieber (1983), De Zanche
237.0 et al. (1995), Brack et al. (1996),
and Zühlke et al. (2003). Chro-
238.0 nostratigraphy: U–Pb ages of
Mundil et al. (2003). Absolute
Curionii 239.0 ages (sampling horizons) were
239.28
UTF Curionii
tied to the cyclostratigraphic
239.83
239.99 240.6 UTF 240.0 data of this study.
Curionii
(241.2
-0.6)
240.34 Remarks.—Extrapolated ages
240.76 241.5
UTF +0.4/ UCF 240.9 UCF 240.99 241.0 shown in columns E1 to E3 are
241.08 -0.6 Sece- Sece-
(LAT-32)
UCF 241.2 241.56
(241.5
-0.6) densis 241.47 densis based on the assumption that
Sece-
241.80 +0.7/
-0.6 densis 241.9 MTF 241.90 242.0
individual small-scale shal-
MTF MTF (LAT-30)
LCF 242.40
242.6
(241.2
+0.7) 242.40 lowing-upward cycles in the
242.64 ±0.7 LCF 242.71 Latemar represent largely con-
LTF Reitzi
242.86
LTF
242.78
(LAT-31) 243.0
LPF LCF 243.3 Reitzi stant periods (∆tav = 2.68, ∆tmax
243.39 LPF
LTF (242.6
Reitzi ≤ 5.85) and that all beats are
LPF +0.7) 244.0
documented in the cyclic suc-
cession. Spectral analysis were
applied to the complete cyclic
Extrapolated duration Extrapolated duration Extrapolated duration
LCF–UTF LCF–UTF LCF–UTF Dura- Latemar succession (LCF–
tion UTF), on the basis of thickness
Ø 1.88 Ma max. ≤ 4.10 Ma 2.95 Ma
of accommodation cycles. The
duration of the Latemar cyclic series has been constrained by cyclostratigraphic, biostratigraphic, and chronostratigraphic data
as well as spectral-analysis data.

Columns E1–E3 compare different sets of constraints used in this study in order to account for maximum cumulative error intervals
in the resulting forcing model.

E1: maximum number of cycles in the Latemar cyclic series (701); average ages (Mundil et al. 2003); biostratigraphic data (Fig. 2,
Columns E1–E2); average period of the basic shallowing-upward microcycle (∆t = 2.68 ky).

E2: maximum number of cycles in the Latemar cyclic series (701); maximum uncertainties in ages (Mundil et al., 2003); biostrati-
graphic data (Fig. 2, columns E1–E2), maximum period of the basic shallowing-upward microcycle (∆t = 5.85 ky).

E3: maximum number of cycles in the Latemar cyclic series (701), spectral analysis (Fig. 12B); ages for LAT-30 to LAT-32 (Mundil et
al., 2003) were shifted within their respective error intervals; biostratigraphic data (Fig. 2, Columns E1–E2); most probable
duration of the basic shallowing-upward microcycle (∆t = 4.2 ky; Figure 12B).
198 RAINER ZÜHLKE

40
Cumulative
deviation
40 from mean 35
cycle thickness
[m]
35 30
Number of
Cycles [-]
30 25

25 20

20 15

15 10

10 5

5 0

0 -5

-5 -10

-10 -15

maximum number of cycles


minimum number of cycles

-15 -20

-20 -25

-25 -30
LCF MTF UCF UTF
-30
1 51 101 151 201 251 301 351 401 451 501 551 601 651 701

FIG. 8.—Raw thicknesses of accommodation cycles in the Latemartürme reference section (Chart 1, Column A3). The Fischer plot
shows cumulative departures from mean cycle thickness versus cycle number (Fischer, 1964; Read and Goldhammer, 1988; Sadler
et al. 1993). The upper curve indicates maximum number of cycles (701), and the lower curve, the minimum numbers of cycles
(619). Dotted lines show identical cyclostratigraphic levels. Thick gray lines represent fourth-order polynomial trendlines. Fischer
plots provide a first indication for higher-order cycle stacking patterns.

reference section. Previous spectral analyses include only the period of individual microcycles (∆tav = 2.68 ky, ∆tmax ≤ 5.85 ky).
uppermost MTF and UCF (Hinnov and Goldhammer, 1991; Spectral analyses were performed with Mathematica® (Wol-
Preto et al., 2001). In this study, spectral analysis is based fram, 1996) including the Time Series Pack (Yu, 1995), and
exclusively on accommodation cycles, which built up to sea AnalySeries© (Paillard et al., 1996).
level at the end of each cycle. Shallowing-upward successions Blackman–Tukey Spectral Analysis.—The Blackman–Tukey
within subtidal depths have not been included, because un- method (Blackman and Tukey, 1958) is the classical method of
known unfilled accommodation space from the previous spectral analysis. The algorithm computes first the autocovariance
microcycle may have existed at the beginning of a new microcycle of the data, then applies a window, and finally Fourier-trans-
starting in subtidal depths. Spectral-analysis methods include forms it to compute the spectrum. According to Paillard et al.
(1) periodogram, (2) maximum entropy, (3) Blackman–Tukey, (1996) it is a very robust method, unlikely to present spurious
(4) multi-taper spectral analysis, and (5) multi-taper harmonic spectral features. The main drawback is its poor resolution in the
analysis. Special focus is placed on Blackman–Tukey, multi- spectral domain; sharp features are considerably smoothed.
taper spectral, and harmonic analyses. The raw thicknesses of Blackman–Tukey analysis of the complete Latemar cyclic series
accommodation cycles (Figs. 8, 9) were subject to three transfor- (LCF–UTF) reveals six clusters of frequencies with moderate to
mations (Fig. 10): (1) linear trend removed, (2) unit variance strong power (Fig. 11A). The distribution of clusters varies over
normalized (even sampling), and (3) Box–Cox transformation. the cyclic succession:
For spectral analysis the cyclic series was subdivided into seven
spectral windows with 50% overlap each to the previous and • Average ratio of 1:3.7—prominent in the LCF to upper MTF,
next spectral window. Spectral windows include between 125 gradually disappears in the UCF to UTF
and 220 accommodation cycles each and therefore represent • Average ratio of 1:5.9—significant in the UCF to lower UTF.
sufficiently large, reliable data windows for spectral analyses. Both ratios are largely lacking in the LCF. They partly appear
All power and amplitude spectra were thickness and time as a double peak with an average ratio of 1:5.2 (upper MTF to
calibrated, constrained by the maximum numbers of cycles in lower UCF, upper UCF)
the Latemartürme section and the biochronostratigraphic data • Average ratios of 1:9.1 and 1:19.4—strong spectral power in
presented above. The time calibrations indicated in Figures the MTF to UCF. Both ratios are completely lacking in the LCF
11A–B and 12A consider both the average and the maximum and middle to upper UTF.
INTEGRATED CYCLOSTRATIGRAPHY—THE LATEMAR (MIDDLE TRIASSIC, ITALY) 199

Cumulative
deviation
from mean
cycle thickness
[m]

Number of
Cycles [-]
5 Spiz de la Cagnota
0
-5
Cumulative deviation from mean cycle thickness [m]

5 Valle de la Vecia
0
-5

5 Latemartürme
0
-5

5 Cima Forcellone
0
-5
5 Cima Valsorda
0
-5

LCF MTF UCF UTF


Number
of Cycles 1 51 101 151 201 251 301 351 401 451 501 551 601 651 701

FIG. 9.—Raw thickness of accommodation cycles in the five vertical sections of this study (Chart 1, Columns A3–A7) as indicated by
Fischer plots (cf. Figure 8). Thick black lines indicate the cyclic succession preserved in each section. Thin black lines indicate cycles
projected from the Latemartürme reference section.

• Average ratio of 1:99.2—low resolution; the cluster shows • Average ratio of 1:24.7—elevated to high power in the MTF to
considerable variations over the spectral windows. UCF. The distribution of this ratio is similar to that of the
average ratio of 10.1.
Multi-Taper Spectral Analysis.— • Average ratio of 1:115.2—low resolution as in Blackman–Tukey
analysis. The variation of frequencies is comparatively low.
The multi-taper method (Thomson, 1982) offers high resolu-
tion and statistical estimates that are independent of the spectral Multi-Taper Harmonic Analysis.—
power (small-amplitude oscillations may have a high significance
level; Yiou et al., 1994). Both multi-taper spectral analysis and In contrast to Blackman–Tukey analysis and multi-taper spec-
harmonic analysis were applied to the complete Latemar cyclic tral analysis, multi-taper harmonic analysis considers exclu-
series (LCF–UTF). Multi-taper spectral analysis reveals six clusters sively statistical confidence levels (F-Test), which are indepen-
of frequencies with moderate to strong power (Fig. 11B). The dent of spectral power (Yiou et al., 1994). It allows identifying
variation in the distribution of clusters through the cyclic succes- low-amplitude peaks with high significance or high-amplitude
sion is similar to that revealed by the Blackman–Tukey analysis: peaks with low significance. F-test results for individual ampli-
tude peaks may vary between high and highest significance,
• Average ratio of 1:3.8—prominent in the LCF to middle MTF. depending on the taper type and the moving window applied.
The ratio disappears completely in the middle UCF to middle Multi-taper harmonic analysis has the advantage of high resolu-
UTF and reappears with moderate power in the middle to tion even in series with high background noise. Background
upper UTF. noise is potentially caused by changes in sedimentation rates over
• Average ratio of 1:5.0—especially significant in the UCF to the interval to be analyzed or by (moderately) nonstationary
lower UTF. The ratio is subdued or absent in the LCF to cycle periods, especially in the high-frequency range. Multi-taper
middle MTF and the middle to upper UTF. harmonic analysis of the complete Latemar cyclic series (LCF–
• Average ratio of 1:6.2—only moderate power and without UTF, Figure 12A) reveals seven clusters of frequencies with high
any clear distribution over the cyclic series. significance (> 0.9) to highest significance (1/(1-n), > 0.994). The
• Average ratio of 1:10.1—shows strong power in the upper variation in the distribution of clusters over the cyclic succession
LCF to UCF. The ratio is lacking in the lower LCF and the is similar to that revealed by Blackman–Tukey and multi-taper
middle to upper UTF. harmonic analysis:
200 RAINER ZÜHLKE

8.00 transformed
individual micro-
5.50
cycle thickness
5.00 [m]

4.50 cycle
number [n]
4.00
linear trend (0)
3.50
3.00
2.50
2.00
1.50
1.00
0.50
0.00
-0.50
-1.00
-1.50
0 50 100 150 200 250 300 350 400 450 500 550 600 650 700

LCF MTF UCF UTF


SW1 SW3 SW5 SW7
SW2 SW4 SW6

FIG. 10.—Basic dataset for spectral analyses in the Latemartürme reference section. The graph plots transformed individual
accommodation cycle thickness (Chart 1, Column A5) versus cycle number. Raw data were subject to the following transforma-
tions (1) linear trend removed; (2) unit variance normalized (even sampling); (3) Box-Cox transformation (not included in this
graph). SW indicates sliding windows used in spectral analyses (see Figures 11 and 12).

• Average ratios of 1:2.2 and 1:2.7—occur only in multi-taper • ratios indicative of orbital forcing in the Triassic (Berger and
harmonic analysis but not in the other four analysis methods. Loutre 1994). With the short precession period set to 1, ratios
The ratios may be produced by the different statistical null of 1:1.9 indicate long precession, 1:2.01 short obliquity, 1:2.54
hypothesis of the method. long obliquity, 1:5.63 short eccentricity, and 1:22.5 long eccen-
• Average ratio of 1:3.7—highly significant between the upper tricity (Table 2, Column G).
LCF to UTF. The ratio is especially strong in the MTF. • time intervals indicative of orbital forcing in the Triassic
• Average ratio of 1:5.2—moderate significance; occurs through- (Berger and Loutre, 1994). Intervals include 17.78 ky and 21.23
out the cyclic series. ky for short and long precession, 35.73 ky and 45.24 ky for
• Average ratio of 1:6.8—occurs only in multi-taper harmonic short and long obliquity, and 100 and 400 ky for long eccen-
analysis but not in the other four analysis methods. The tricity (Table 2, Column J)
significance is moderate in the LCF to MTF and upper UCF to • indicative ratios and time intervals must be compatible with
lower UTF. the chronobiostratigraphic data, i.e., must be applicable with
• Average ratio of 1:9.9—significant between the middle LCF ∆tmax ≤ 5.85 ky.
to UTF. Highest significance occurs in the lower to middle • the first three requirements must be met by significant to
MTF and the UCF. highly significant amplitude peaks in multi-taper harmonic
• Average ratio of 1:19.5—significant between the middle LCF analysis.
and lower UTF. It includes a double peak between the upper
MTF and the UCF. In contrast to the approach of Preto et al. (2001), which was
• Average ratio of 1:105—significant in the MTF to UTF. High- based exclusively on spectral analyses with a Milankovitch-
significance points occur in the UCF to lower UTF. model-dependant calibration and which considered only a mi-
nor part of the Latemar cyclic succession, this approach consid-
Results of Spectral Analysis.— ers all available constraints: (1) the complete 458-m-thick cyclic
succession in the Latemar with 701 accommodation cycles; (2)
All five analysis methods indicate five ratios with strong relative ages and durations indicated by biostratigraphy; (3)
power and/or high significance over the largest part of the absolute ages derived from chronostratigraphy, which were
Latemar cyclic series with average ratios of: (1) 1:3.7 to 1:3.8; (2) extrapolated to the complete LCF–UTF succession; (4) spectral
1:4.7 to 1:5.2; (3) 1:9.1 to 1:10.1; (4) 1:19.4 to 1:24.7; (5) 1:99.2 to 1:105. analyses with a model-independent time calibration of
Basic shallowing-upward microcycles in the Latemar cyclic suc- microcycles.
cession cannot be reconciled with precession forcing but repre- The approach uses a goal seek script that includes the four
sent sub-Milankovitch control. In order to check whether Mi- requirements stated above as nested conditions. The script runs
lankovitch forcing is additionally present in the larger-scale cycle recursively on all seven sliding windows with steps of 0.1 ky from
bundlings of the Latemar cyclic series, the results of multi-taper ∆t > 0 ky to ∆tmax ≤ 5.85 ky. Bandwidths of Milankovitch signals
harmonic analysis were tested for a mathematical solution that include an error interval of ≤ ± 6% to account for geological
meets all of the following four requirements: distortions (e.g., hiatuses, varying sedimentation rates; cf. Stage,
INTEGRATED CYCLOSTRATIGRAPHY—THE LATEMAR (MIDDLE TRIASSIC, ITALY) 201

1999). In fact there is a single mathematical solution that meets all relative. This indicates a simultaneously aggrading sheet model
four requirements stated above (Fig. 12B). Ratios and time inter- sensu Pratt et al. (1992), which was described for ancient platform
vals indicative of orbital forcing in the Milankovitch band exist in settings by, e.g., Koerschner and Read (1989). Other peritidal
the Latemar cyclic series with high and very high significance platform models like tidal wedges or multiple tidal-flat islands
points when ∆t = 4.2 ky. The period of ∆t = 4.2 ky for each sub- include meter-scale successions that are laterally discontinuous
Milankovitch shallowing-upward microcycle is well within the and cannot be traced or correlated over kilometer distances
range of the average to maximum period indicated by the chrono- (except for their basal disconformity). The tidal-wedge model
stratigraphy and biostratigraphy (∆t av = 2.68 ky, ∆t max ≤ 5.85 ky). necessitates pronounced lateral progradation from some sort of
Provided that isotopic ages reflect true ages of deposition, and nucleus. Progradation typically shows simple or staggered offlap
on the basis of solution above, at least five different hierarchic with progradation jumps between barriers (Hardie and Shinn,
cyclicities exist in the Latemar cyclic series (Table 2). Ratios are 1986). These features are absent from the Latemar platform
expressed both with the sub-Milankovitch basic microcycle of 4.2 interior. The tidal-flat-island model requires complex lateral and
ky set to 1 (Table 2, Columns C–D) and the higher-order short- vertical shifting of depositional facies in response to local varia-
precession cycle of 17.8 ky set to 1 (Table 2, Columns E–F): tions in topography, water energy, and currents. Because indi-
vidual microcycles in the Latemar platform interior show moder-
• sub-Milankovitch cyclicity: basic shallowing-upward ate lateral variations in lithofacies, it cannot be completely ruled
microcycle, average microcycle thickness 0.63 m (e.g. in UCF), out that minor, low-angle progradation took place within single
period ∆t = 4.2 ky. The basic shallowing-upward microcycle microcycles. However, no progradational features including sev-
was previously interpreted as a precessional cycle of ∆t ~ 20 ky eral microcycles, macrocycles, or larger-scale cycle bundles exist.
(Goldhammer et al., 1993). Preto et al. (2001) attributed a Models of internal control (autocyclicity models) assume
periodicity of ∆t = 17.6–21.7 ky (precession forcing) to the localized changes in carbonate accumulation rates as the driving
shortest cyclic signal in their rank series. force for shallowing-upward cycles. Autocyclic processes may
• condensed precession cycle or higher-order sub-Milanko- generate successions that are similar to those that are inferred to
vitch cycle: stacking pattern of 1:3.2 to 1:4.0 microcycles, have formed by eustatic, orbitally driven sea-level changes. The
average cycle thickness 2.27 m (e.g., in UCF), ∆tmax = 13.6–16.7 original model of Ginsburg (1971) was based essentially on a
ky. prograding-wedge model. On platform tops with simultaneous
• short and long precessional cycle: stacking patterns of 1:4.3 vertical aggradation, an autocyclic model implies that sediment
to 1:5.1 microcycles, average cycle thickness 2.96 m (e.g., in production rates in subtidal and peritidal to intertidal depths are
UCF), ∆tmax = 17.1–21.5 ky. This stacking pattern was previ- sufficiently high to aggrade to sea level. This is difficult to achieve
ously interpreted as an eccentricity cycle of ∆t ~ 100 ky with an exclusively autocyclic model because the carbonate
(Goldhammer et al., 1993) or ∆t = 98 ky (Preto et al., 2001). factory is shut down progressively with increasing extension of
• short and long obliquity cycle: stacking patterns of 1:8.5 to intertidal areas. On small platforms with an only slightly elevated
1:11.4 microcycles, stacking patterns of 1:2.0 to 1:2.7 short- topographic margin like the Latemar, wave activity occurs over
precession cycles, average thickness 6.24 m (e.g., in UCF), the whole platform top and inhibits cyclic aggradation to sea level
period ∆tmax = 35.5–48.0 ky. exclusively driven by autocyclic processes.
• short-eccentricity cycle: stacking patterns of 1:22.7 to 1:25.2 Models of external control (allocyclic models) include struc-
microcycles, stacking patterns of 1:5.4 to 1:6.0 precession tural processes or low-amplitude, high-frequency sea-level
cycles, average cycle thickness 15.10 m (e.g., in UCF), period changes as the dominant driving force for rhythmic deposition.
∆tmax = 95.4–105.9 ky. Structural processes that operate on a 10–100 ky scale include
• long-eccentricity cycle?: stacking pattern of 1:64.3 to 1:147.0 the inflation of magma chambers, accompanied by doming of a
microcycles (average 1:105.7), stacking pattern of 1:15.2 to few tens of meters of kilometers across and eruption followed by
1:34.7 precession cycles, average cycle thickness 66.6 m (e.g., deflation (Sloan and Williams, 1991). The Latemar is situated in
in UCF), period ∆tmax = 270.1–617.3 ky (∆tav = 444 ky). The the vicinity of the magmatic center of Predazzo and Monzoni
existence of the long-eccentricity signal remains speculative (Bosellini et al., 1982). However, significant magmatic and volca-
because ratios show considerable variations. nic activity occurred only in the early late Ladinian (Longobardian,
upper Gredleri to Archelaus Zone) at an age of 237.2 +0.4/-1.0 Ma
High-resolution cyclostratigraphic, biostratigraphic, and (Mundil et al. 1996a). It postdated the deposition of the preserved
chronostratigraphic data in combination with the results of spec- Latemar carbonate platform interior for at least 2.3 My and 4.6 My
tral analyses provide an integrated set of constraints on the at maximum. High-frequency structural processes, e.g., stick-slip
duration of the Latemar cyclic succession (Fig. 7, Column E3). The faulting at the platform margin (Cisne, 1986), can be ruled out,
extrapolated duration for the LCF to UTF amounts to 2.95 My. because they depend on lithospheric flexure on a lateral scale of
This equals the Secedensis Zone in the biostratigraphic scheme of > 100 km.
Brack and Rieber (1993) and Brack et al. (1996). For comparison, The internal architecture of the Latemar platform interior
Figure 7, Columns A–D, list the constraints and resulting abso- favors a predominantly allocyclic control of individual micro-
lute durations of the Latemar cyclic series as assumed in previous cycles, by high-frequency sea-level changes in the sub-
studies. Milankovitch bandwidth. Each microcycle reflects a single ex-
cursion of sea level. Low-amplitude, high-frequency sea-level
DISCUSSION changes triggered by orbital forcing via climatic changes pre-
dominantly controlled the large stacking patterns of 1:4.3 to
Platform Architecture, Allocyclicity, and Autocyclicity Models 1:5.1 and higher.

The internal architecture of the Latemar platform interior is Biostratigraphy and Chronostratigraphy
characterized by laterally continuous shallowing-upward
microcycles and thinning-upward macrocycles, the large major- Biostratigraphic information on the ammonoid biozonal level
ity of which are platform-wide and physically traceable or cor- does provide relative estimates of time covered by the Latemar
202 RAINER ZÜHLKE

A B
Blackman–Tukey spectral analysis Multi-taper spectral analysis (Thomson, 1982)
Stacking patterns (ratios, periods; basic microcycle = 1) Stacking patterns (ratios, periods; basic microcycle = 1)

1:60.0–138.3; Ø 1:99.2 double peak 1:102.4–128.0; Ø 1:115.2 1:5.9–6.6; Ø 1:6.2


SW 1:4.4–6.1; Ø 1:5.2
double peak double peak 1:4.4–5.7; Ø 1:5.0
Sliding 1:13.8–25.0; Ø 1:19.4 1:5.8–6.1; Ø 1:5.9 1:15.3–34.1; Ø 1:24.7
Windows 1:4.4–5.1; Ø 1:4.7 1:3.3–4.4; Ø 1:3.8
1:20.1–25.0; Ø 1:22.6 1:15.3–20.1; Ø 1:17.7
1:13.8–1:14.8; Ø 1:14.3 1:26.3–34.1; Ø 1:30.2
see Figs. 1:3.3–4.1; Ø 1:3.7 15.4 frequency, period
2, 3, 5 1:8.8–9.5; Ø 1:9.1 1:7.9–12.3; Ø 1:10.1

8.00 3.50
24.9
SW7 3.00 11.4
26.6 23.8
2.50 32.7 29.7 12.2
6.00 15.0 13.6 10.9
Top 30.3 27.1 24.8
13.9 12.4 11.0 599.0
468.40 m 2.00 274.4
Top UTF 35.4
4.00 16.2 1.50
Base
354.45 m 1.00
upper UCF 2.00
0.50
182 cycles
0.00 0.00

8.00 3.50
SW6 29.1
3.00 105.1 13.3
48.1 25.7
82.1 2.50 199.7 11.8
Top 6.00 27.4
91.5 95.1 32.4
37.6 12.5 43.6
411.45 m 748.8 14.8
351.0 2.00 343.1 36.8
middle UTF 160.8 128.7 46.4
16.8
4.00 59.0 33.9
23.4 1.50 21.3
Base 15.5
10.7
297.50 m
middle UCF 1.00
2.00

167 cycles 0.50


0.00 0.00
3.50 30.3 28.3 27.4
SW5 8.00 13.9 12.9 12.5
3.00 176.2
28.6 80.7
13.1 25.6
117.8 2.50 101.5 67.3
Top 6.00 353.1 46.5 30.8 47.5
11.7
54.0
354.45 m 161.8 21.8
86.3 55.5 2.00
upper UCF 39.5 25.4
4.00 23.0
Base 10.5
1.50
240.50 m
lower UCF 1.00
2.00
167 cycles 0.50
0.00 0.00
138.7 3.50
SW4 8.00
63.6
748.6
3.00 343.1 110.9
30.0 50.8 56.0 31.7
13.7 25.6 14.5 28.8
Top 6.00 54.6 2.50 153.6 13.2
809.2 25.0 70.4
297.50 m 60.5
370.7 2.00 98.2 27.7 25.8
middle UCF 45.0 11.8 23.3
4.00 80.9 23.9
34.4 11.0 10.7
Base 37.1 15.8
1.50
183.50 m
upper MTF 1.00
2.00
208 cycles 0.50
0.00 0.00
3.50
SW3 8.00
146.1 52.2
23.9
199.7
56.0
25.6 51.2
66.9 3.00 91.5
72.2 23.5 25.0
82.5 117.5 33.1 11.4
Top 6.00 37.8 2.50 53.8
240.50 m 26.1 89.4 33.5 26.9 23.6
12.0 22.2
2.00 41.0 15.3 12.3 10.8 21.2
lower UCF 464.8 10.2
34.8 36.5 9.7
4.00 212.9
Base 15.9 19.1 1.50 16.7
8.7
126.55 m
lower MTF 1.00
2.00
220 cycles 0.50
0.00 0.00
3.50
SW2 8.00
3.00 21.1
23.4 9.7
2.50 52.1
Top 6.00 51.4 22.1 748.8 58.7 23.9 10.7
183.55 m 23.5 10.1 343.1 26.9
25.5 19.9 2.00 45.4 25.5 19.5
upper MTF 9.1 107.0 11.7 8.9
126.9 11.7 49.0 20.8
4.00 58.1
Base 1.50
69.75 m
middle LCF 1.00
2.00
175 cycles 0.50
0.00 0.00

3.50
8.00 Period microcycle 32.6 Period microcycle
Power (whole spectrum)
Power (whole spectrum)

SW1 ∆tmax = 5.85 ky 3.00 14.9


∆tmax = 5.85 ky
36.1
6.00 ∆tav = 2.68 ky 2.50 16.5 30.9 ∆tav = 2.68 ky
Top 38.4 14.1
33.7 20.3 2.00 17.6 34.4
126.55 m 15.5 24.2 20.9
9.3 15.8 11.1
lower MTF 4.00 9.6
1.50
Base
12.60 m 1.00
2.00
base LCF
0.50
125 cycles
0.00 0.00
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40 0.45 0.50 0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40 0.45 0.50

Frequency (cycles per unit, ∆t = 1) Frequency (cycles per unit, ∆t = 1)

FIG. 11.—A) Blackman–Tukey spectral analysis and B) multi-taper harmonic analysis of the complete cyclic series in the Latemartürme
reference section (LCF–UTF). Color bars indicate ratios of prominent frequencies to the basic microcycle. Periods are based on
the average and maximum cycle periods of ∆tav = 2.68 ky and ∆tmax ≤ 5.85 ky calculated from biochronostratigraphic constraints
and the maximum number of cycles bracketed by ash layers LAT-31 and LAT-32 (Fig. 7, Columns E1, E2; Chart 1, Column A5).
Legend: SW1-7 = sliding windows 1–7 (Fig. 10).
INTEGRATED CYCLOSTRATIGRAPHY—THE LATEMAR (MIDDLE TRIASSIC, ITALY) 203

A B
Multi-taper harmonic analysis (Thomson, 1982) Multi-taper harmonic analysis (Thomson, 1982)
Stacking patterns (ratios, periods; basic microcycle = 1) Milankovitch forcing model (basic microcycle ∆t = 4.2 ky)

400 ky Eccen- 21.3 ky


1:64–146.3; Ø 1:105 1:3.2–4.1; Ø 1:3.7 100 ky tricity Ø 19.5 ky Precession SW
1:13.5–25.6; Ø 1:19.5 1:2.4–3.0; Ø 1:2.7 17.7 ky
Obliquity Sliding
1:8.5–11.4; Ø 1:9.9 1:2.0–2.3; Ø 1:2.2 16.0 ky condensed Windows
45.3 ky precession or
1:6.1–7.4; Ø 1:6.8 15.4 s ≥ 0.994 (1-1/N), period Ø 14.9 ky
Ø 40.5 ky higher-order see Figs.
1:4.3–6.0; Ø 1:5.2 15.4 s ≥ 0.9, period 13.8 ky sub-Milanko- 2, 3, 5
35.7 ky vitch cyclicity

0.50 0.50
29.8 23.4
SW7
0.40 13.7 10.7 0.40 21.5 16.9
855.8 119.8 36.5 13.8
16.7 18.9 26.3
392.0 54.9 6.3 617.3 86.4 10.0
Top
0.30 8.7 16.6 14.2 0.30 13.6
66.6
7.6 6.5 10.2 468.40 m
30.5 12.0
15.5 48.0 Top UTF
20.7 7.1 11.2
0.20 9.5 12.0 0.20 14.9 Base
5.5 8.6
354.45 m
0.10 0.10 upper UCF
182 cycles
0.00 0.00
0.50 82.1 0.50
499.2
228.7 37.6
357.5 58.8 SW6
0.40 54.5 34.8 0.40
127.5 24.9 41.0 16.0 24.9
58.4 18.8 23.7 22.6
91.3 29.4 Top
16.8
0.30 49.5 10.8 10.4 19.3 7.7 0.30 17.0 411.45 m
22.7 35.5 12.0
8.8 14.1 39.0 16.2 13.8 middle UTF
6.5 10.1
12.4 Base
0.20 15.1 5.7 0.20 8.9
6.9 10.8 297.50 m
middle UCF
0.10 0.10
167 cycles
0.00 0.00
0.50 374.4 55.5 27.7 0.50 270.1 20.0
171.5
85.6
25.4 29.8
13.7
12.7 40.0
21.5
SW5
0.40 142.8 39.2 0.40 61.7
65.3 23.1 102.9
10.6
Top
43.4 16.7
0.30 0.30 354.45 m
19.9 19.5 17.4 31.3
8.9 8.0 12.5
upper UCF
15.2 12.8 14.1
14.0 9.3 Base
7.0 5.9 11.0 10.1
0.20 6.4 0.20
240.50 m
lower UCF
0.10 0.10
167 cycles
0.00 0.00 0.10 0.20

0.50 0.50
855.8
149.8
68.6 53.5 105.9
SW4
392.0 24.5 605.1 37.8
0.40 30.4 0.40
78.8 13.9 21.5 Top
34.8 24.2 19.3
36.1 11.1 55.7
0.30 16.0 8.9 0.30 24.6 17.1 297.50 m
12.7 13.7 middle UCF
18.8 14.3 5.8 8.8
22.5 20.1 8.6 6.5 15.9 10.1
10.3 9.2 13.3 Base
0.20 16.8 0.20 14.2
12.7
7.7 15.7 11.9 183.50 m
7.2 11.1
upper MTF
0.10 0.10
208 cycles
0.00 0.00
0.50 0.50
52.5 22.4 SW3
83.2 24.1 10.2 16.0
0.40 142.6 38.1
26.2 0.40 37.5
12.0 59.4 18.7
65.3 19.1 101.9 Top
8.7 13.6
34.2 16.2 240.50 m
0.30 15.7 7.4 0.30 24.5
30.1
lower UCF
17.1 12.0
13.8 7.8 15.8 14.7 21.5
40.8 5.5 12.2 11.3 8.6 Base
0.20 18.7
7.3
6.7
13.5 0.20 11.6 10.5
6.2 29.1 9.6 126.55 m
lower MTF
0.10 0.10
220 cycles
0.00 0.00
0.50 21.9 0.50
50.3
23.1
10.0
36.1
15.7 SW2
748.8
0.40 343.1 0.40 536.4
133.1 25.3 Top
61.0 62.4 35.7 16.1 95.3
28.6 11.6 18.1
0.30 16.3 7.4 11.7 0.30 44.7 25.5 11.6 183.55 m
42.5 5.4 upper MTF
19.5 30.4 8.4
15.2
7.0 12.0 10.9 Base
0.20 31.7 5.5 0.20 8.6
14.5 22.7 69.75 m
middle LCF
0.10 0.10
175 cycles
0.00 0.00

0.50 0.50
35.7 Period microcycle Period microcycle
0.40
16.3
25.4 ∆tmax = 5.85 ky 0.40
25.5
∆t = 4.2 ky SW1
43.4 11.6 16.4 18.2
19.9 7.5 ∆tav = 2.68 ky 31.1 11.7
Amplitude

Amplitude

0.30 0.30 Top


11.7 126.55 m
5.4 8.4 lower MTF
0.20 0.20
12.7 Base
5.8 9.1
12.60 m
0.10 0.10 base LCF

0.00 0.00 125 cycles

0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40 0.45 0.50 0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40 0.45 0.50

Frequency (cycles per unit, ∆t = 1) Frequency (cycles per unit, ∆t = 1)

FIG. 12.—Multi-tapered harmonic analyses of the complete cyclic series in the Latemartürme reference section (LCF–UTF). A) Color
bars indicate ratios of (highly) significant amplitudes compared to the basic microcycle. Periods are based on the average and
maximum cycle periods of ∆tav = 2.68 ky and ∆tmax ≤ 5.85 ky calculated from biochronostratigraphic constraints and the maximum
number of cycles bracketed by ash layers LAT-31 and LAT-32 (Fig. 5, Columns E1 and E2; Figure 4; Chart 1, Column A5). B)
Milankovitch forcing model of the cyclic series based on ∆t = 4.2 ky (Fig. 5, Column E3; Table 2, Columns A–F). Color bars indicate
condensed precession or higher-order sub-Milankovitch cyclicity (red, 13.8–16.0 ky), short and long precession (blue, 17.7 ky, 21.3
ky); short and long obliquity (green, 35.7 ky, 45.3 ky), short eccentricity (brown, 100 ky) and questionable long eccentricity (gray,
400 ky). Legend: s = significance point, N = sampling number, SW1–7 = sliding windows 1–7 (Fig. 8).
204 RAINER ZÜHLKE

TABLE 2.—Sub-Milankovitch and Milankovitch cycles in the Latemar platform interior compared to Triassic orbital parameters.
The table lists statistically significant periods (column A), cyclicities (B), ratios (C), and average ratios (D) of cycle bundles
in the Latemar when the basic shallowing upward microcycle is set to 1 and ∆t = 4.2 ky based on biochronostratigraphic data
and spectral analyses (see Figure 5, Column E3). Columns G–J list average ratios, ratios, cyclicities, and periods of Triassic orbital
forcing (Berger and Loutre, 1994; interpolation between 270 and 72 Ma). Average ratios (G) and ratios (H) are based on the short
precession set to 1 and ∆t = 17.78 ky. Given the solution presented above, the duration of the 1:4.7 stacking pattern (see Column
D) in the Latemar closely matches the precessional period in the Triassic. Setting this precessional signal to 1 and ∆t = 17.78 ky,
periods (A), ratios (E), and average ratios (F) of the large-scale cycle bundles in the Latemar closely match periods (J), average
ratios (G), and ratios (H) of Triassic orbital parameters. Columns F and G indicate the consistent average ratios of orbitally forced
cycle bundles in the Latemar and of Triassic orbital parameters (in boldface type).

This study - Latemar carbonate platform interior Orbital parameters - Triassic

(A) (B) (C) (D) (E) (F) (G) (H) (I) (J)

Dt min. Ratio ø Ratio Ratio ø Ratio ø Ratio Ratio


Dt
Dt max. Cyclicity Cyclicity
microcycle = 1 precession cycle = 1 precession cycle = 1 [ky]
[ky] Dt = 4.2 ky Dt = 17.78 ky Dt = 17.78 ky

basic microcycle
4.2 1 : 1.00 1 : 1.00 – – – – – –
sub-Milankovitch

cond. precession
13.63 1 : 3.24
or higher order 1 : 3.61 – – – – – –
16.68 1 : 3.98
sub-Milankovitch

18.10 1 : 4.31 1 : 1.02 1 : 1.00 short precession 17.78


precession 1 : 4.72 1 : 1.11 1 : 1.10
21.50 1 : 5.12 1 : 1.21 1 : 1.19 long precession 21.23

35.46 1 : 8.45 1 : 2.00 1 : 2.01 short obliquity 35.73


obliquity 1 : 9.94 1 : 2.35 1 : 2.28
48.01 1 : 11.43 1 : 2.70 1 : 2.54 long obliquity 45.24

95.35 1 : 22.69 1 : 5.36


short eccentricity 1 : 23.95 1 : 5.66 1 : 5.63 1 : 5.63 short eccentricity 100.00
105.89 1 : 25.21 1 : 5.96

270.08 1 : 64.31 1 : 15.19


long eccentricity
?
1 : 105.65 1 : 24.95 1 : 22.50 1 : 22.50 long eccentricity 400.00
617.32 1 : 146.98 1 : 34.71

cyclic series. It has been argued that this estimate depends on Therefore the relative biostratigraphic ranges of the Latemar
whether it is based on the biostratigraphic dataset of Brack and cyclic succession are almost identical in the two biostratigraphic
Rieber (1993), Brack et al. (1996), and Zühlke et al. (2000), or on the schemes—little more than a single biozone: upper Reitzi Zone to
dataset of de Zanche et al. (1995). Because the boundary between uppermost Secedensis Zone or (?) lowermost Curionii Zone (Brack
the Middle Triassic stages of the Anisian and the Ladinian (Illyrian and Rieber, 1993; Brack et al., 1996), upper Hungarites Zone to
to Fassanian substage boundaries) has not yet been formally middle Nevadites Zone (Mietto and Manfrin, 1995, De Zanche et
defined (see Brack and Rieber, 1994), all biostratigraphic ages al., 1995). This consistent estimate is independent of biozonal or
have to refer to the ammonoid biozone level. It is necessary to subzonal definitions or different levels proposed for the Anisian—
correlate the biozones in the two schemes on the basis of indica- Ladinian boundary.
tive faunas (Fig. 2, Columns E1, E2). Internal plausibility checks on the genetic cyclostratigraphic
The ammonoid fauna AF-1, including Aploceras avisianum, models for the Latemar relate (1) the durations of the Latemar
Parakellnerites rothpletzi, Flexoptychites noricus, and Latemarites cyclic series as proposed or implied by the different cyclostrati-
latemarensis, occurs in the uppermost part of the LPF. In the graphic models; (2) biostratigraphic data for the Triassic system;
scheme of Brack and Rieber (1993) and Brack et al. (1996) this and (3) chronostratigraphic data for the Triassic System. According
fauna indicates upper Reitzi Zone. In the scheme of Mietto and to the time scale of Gradstein et al. (1994, 1995) and the
Manfrin (1995) and De Zanche (1995) it indicates the middle to biochronostratigraphic chart of Hardenbol et al. (1998), the Per-
upper Avisianum subzone/upper Hungarites Zone. Chieseiceras mian–Triassic boundary has an age of 248.2 ± 4.8 My, and the
sp., as included in the ammonoid fauna AF-8, at the top of the UTF Triassic–Jurassic boundary an age of 205.7 ± 4.0 My. The duration
indicates the uppermost Secedensis to (?) lowermost Curionii Zone of the Triassic is 42.5 My (based on average ages) and ≥ 33.7 My or
in the scheme of Brack and Rieber (1993) and Brack et al. (1996). ≤ 51.3 My (based on maximum uncertainties). The Triassic series in
In the scheme of Mietto and Manfrin (1995) and De Zanche et al. the Alpine–Tethyan domain includes 31–34 ammonoid biozones.
(1995) it indicates the Chiesense subzone to upper Nevadites Zone. The Latemar cyclic series (LCF–UTF) covers about a single am-
Both schemes define the base of their Curionii Zone, Curionii monoid biozone. This Middle Triassic biozone is calibrated by the
subzone, respectively, by the first occurrence of Eoprotrachyceras absolute durations that the different cyclostratigraphic models
curionii. The top of the UTF (AF-8) is slightly below or above this require for the cyclic series (Fig. 7): (1) ≤ 14 My, but at least 12 My
well defined boundary. according to Hinnov and Goldhammer (1991) and Goldhammer
INTEGRATED CYCLOSTRATIGRAPHY—THE LATEMAR (MIDDLE TRIASSIC, ITALY) 205

et al. (1993); (2) 3.20–3.95 My for the uppermost MTF to upper on cumulative errors and spectral-analysis data, ∆t ≤ 4.2 ky) of the
UCF according to Preto et al. (2001), which implies at least 10 My short-precession period in the Triassic. Paleozoic and Mesozoic
for the LCF–UTF; (3) 1.88–4.10 My (cyclostratigraphic, biostrati- sub-Milankovitch periodicities have been reported, e.g., by Bond
graphic, and chronostratigraphic constraints) and 2.95 My (in et al. (1993) for the Cambrian and Yang and Kominz (1999) for the
combination with spectral analysis data) according to this study. Pennsylvanian. However, these periodicities were inferred ex-
On the basis of the forcing models of Goldhammer et al. (1993) clusively from spectra that were calibrated to match the Milanko-
and Preto et al. (2001) the complete Latemar cyclic series or the vitch spectrum. No biostratigraphic or chronostratigraphic ages
single Middle Triassic Latemar biozone represents min.-max. 19– existed, which could have been used to perform a model-inde-
36% (including all error intervals) or an average of 24–28% of the pendent calibration. Yang and Kominz (1999) therefore assumed
complete Triassic duration. The Secedensis Zone or the Nevadites that the sub-Milankovitch peaks in their spectra represented
Zone in the schemes of Brack and Rieber (1993) and Mietto and random components. In contrast, the existence of sub-Milanko-
Manfrin (1995) was min.-max. 7.0–8.3 times as long as the average vitch periodicities in the Latemar cyclic succession is substanti-
duration of the other 30–33 Triassic biozones. ated both by high statistical significance in spectral analyses and
On the basis of the forcing model of this study, the Latemar by high-resolution biochronostratigraphic ages.
cyclic series or the single Middle Triassic Latemar biozone repre- Quaternary sub-Milankovitch cyclicities with periods of 1–10
sents min.-max. 4–12%, probably 7%, of the complete Triassic ky have been studied in detail since the late 1980s. They include
duration (based on a total duration of 2.95 My of the Latemar the Dansgaard–Oeschger (D–O) cycles (Broecker and Denton,
platform interior; see Fig. 7, column E3). The Secedensis Zone or 1989), Heinrich events (Heinrich, 1988; Broecker et al., 1992) and
the Nevadites Zone in the schemes of Brack and Rieber (1993) and Bond cycles (Bond et al., 1992). D–O cycles were initially detected
Mietto and Manfrin (1995) was min.-max. 1.3–2.9 times, probably in cores from the Greenland ice shield (Broecker and Denton,
2.1 times, as long as the average duration of the other 30–33 1989). Subsequent research has proved that D–O cycles are wide-
Triassic biozones. spread in marine sediments on northern Atlantic continental
Mundil et al. (2003, their table 1) document isotopic data shelves (Rasmussen et al. 1998). Furthermore, D–O cycles have
analyzed by single-crystal techniques for the three ash-fall hori- also been found in deep-sea sediments off Namibia and in cores
zons LAT-30 to LAT-32 in detail and discuss potential and from the Antarctic ice sheet. D–O cycles, like other sub-
existing uncertainties in 206Pb/238U dating. Uncertainties are Milankovitch cycles, are strongly asymmetric, with periods be-
caused by secondary Pb loss, the statistical approaches applied to tween 1.5 to 5.0 ky over the last 50 ky. Nonstationary periodicity
calculate mean ages and maximum errors, and the presence of is probably caused by stochastic resonance, i.e., the combination
inherited cores in zircons. Mundil et al. (2003) present clearly of a dominating sinusoidal fluctuation in ocean–atmosphere
identifiable age clusters for LAT-30 to LAT-32 and conservative circulation with another subordinate arbitrary fluctuation. Ac-
criteria for the data selection in isochore plots. cording to existing models, D–O cyclicity occurs on an interhemi-
Radiometric ages for the Latemar, measured by Mundil et al. spheric, global scale. It is controlled by coupled changes in the
(2003), are supported by integrated U–Pb geochronology and ocean–atmosphere circulation, which in turn result in high-fre-
ammonoid biochronology in the Balaton Highlands, Hungary quency climate changes.
(Palfy et al., 2002). Multiple concordant analyses of multi-grain Heinrich events (Heinrich, 1988) are documented in the deep-
zircon fractions in volcanic ash layers in the Felsöors section yield sea record by the cyclic occurrence of levels rich in ice-rafted
ages of 241.1 ± 0.5 Ma to 240.4 ± 0.4 Ma in the Reitzi Zone, and 238.7 debris. They are controlled by rhythmic instabilities in the Arctic
± 0.6 Ma in the Gredleri Zone. Palfy et al. (2002) estimate a duration ice shields leading to melting–glaciation pulses. Bond cycles
of less than 5 My for the stratigraphic interval Reitzi Zone to Top (Bond et al., 1992) are saw-toothed cycles of Heinrich events with
Ladinian (cf. Fig. 7, Columns C, E1–E3). The Latemar cyclic series, a 7.2 ky oscillation. Both Heinrich events and Bond cycles are
with 701 basic shallowing-upward microcycles, covers only a inevitably linked to the (sub-)recent wide polar ice sheets. In the
single biozone within this stratigraphic interval of 5 My. The age Triassic, significant polar ice sheets apparently did not exist
data for the Middle Triassic from Hungary provide independent (Wilson et al., 1994).
evidence that microcycles in the Latemar do not represent preces- In general, Quaternary sub-Milankovitch cycles are linked to
sion forcing, but instead sub-Milankovitch forcing. modern icehouse climate conditions, the current plate-tectonic
In summary, the orbital forcing model of Goldhammer et al. configuration, and associated oceanic circulation patterns. There-
(1993) and Preto et al. (2001) cannot be reconciled with the fore, modern sub-Milankovitch cycles like the Dansgaard–
existing high-resolution biochronostratigraphic framework for Oeschger cycles should not be taken as direct analogies for sub-
the Triassic and the Anisian–Ladinian. This framework includes: Milankovitch cycles in the early Mesozoic. The latter occurred
(1) the biostratigraphic dataset existing for the Latemar platform; under relative greenhouse climate conditions and a superconti-
(2) biostratigraphic schemes for the (Middle) Triassic; (3) the nent plate-tectonic configuration with oceanic circulation pat-
isotopic ages according to Mundil et al. (2003); and (4) the terns different than today. However, similarly to the recent
biochronostratigraphic chart of Hardenbol et al. (1998). In con- climate conditions, Triassic climate was characterized by large
trast, the combined sub-Milankovitch and Milankovitch forcing seasonal and latitudinal temperature gradients (Wilson et al.,
model of this study for the Latemar cyclic series is well compat- 1994). Atmospheric general circulation models (AGCMs) of
ible with the high-resolution biochronostratigraphic framework. Scythian and Carnian climate indicate latitudinal surface–air
temperature gradients of up to 60°C between 0° to 60°N. Wide-
Sub-Milankovitch Cyclicity spread and dominant seasonal aridity characterized Triassic
continental interior lowlands. Ranges in seasonal temperature
The new model for the Latemar cyclic series bears important are above 45°C and reach up to 70°, which in the recent icehouse
implications both for early Mesozoic platforms and for carbonate climate is approached only in Central Siberia, Canada, and
platforms during greenhouse climate conditions in general. The Mongolia. Low–latitude areas and marine shelves bordering the
strongest control on platform-interior deposition was exerted by western and northwestern Tethys showed ranges in seasonal
sub-Milankovitch processes with periods of 33% (based on cu- temperature between 10 and 30°C. Strong monsoonal circulation
mulative errors, ∆tmax ≤ 5.85 ky), most probably only 24% (based with very high precipitation reached the western Tethys and
206 RAINER ZÜHLKE

adjacent northern highlands in the summer. Latitudinal tempera- (1991) used the cycle-thickness series included in Goldhammer
ture gradients and strong seasonal climate differences represent (1987, their Appendix 2, 4, L-13; see Chart 1, Column A1). The
an important control on large-scale and strong coupled ocean– definition of microcycles in Goldhammer (1987) was based
atmosphere circulation. The Triassic Period held a very signifi- partly on a model-dependent approach. Where individual shal-
cant potential for fluctuations in coupled ocean–atmosphere lowing-upward cycles could not be recognized because thin-
circulation in the sub-Milankovitch bandwidth. Similarly to the ning-upward megacycles (macrocycles of this study) were con-
recent climate and D–O cycles, the exact link in the Triassic densed, five microcycles were assumed to exist. For instance,
between fluctuations in coupled ocean–atmosphere circulation Goldhammer (1987, App. 4) recognized 6–7 megacycles in the
and sub-Milankovitch sea-level changes is unclear. However, it is stacked tepee zone (middle TF) and schematically attributed 5
well possible that rapid oceanographic and/or atmospheric microcycles to each of them, although he could not recognize or
changes triggered a strong sub-Milankovitch signal in the Middle measure them in the section. In fact, every single megacycle of
Triassic. the 97 megacycles described by Goldhammer (1987) includes
five shallowing-upward microcycles.
Milankovitch Cyclicity Preto et al. (2001) used a four-rank series based on lithofacies
types (Chart 1A, Column A2). Lithofacies types included (1)
Orbital-forcing models for icehouse climate periods (late Ceno- supratidal caliches and soils; (2) supratidal-flat laminated dolo-
zoic, late Paleozoic) are based on the direct coupling between mites; (3) restricted subtidal wackestones; (4) open subtidal pack-
changes in insolation, climate change, waning and waxing of stones to grainstones. The authors stressed that their continu-
polar ice sheets, and sea-level change. The coupling of changes in ously depth-ranked series permitted a detailed, better compari-
insolation and sea level via polar ice sheets cannot be applied to son of the series with a full theoretical Milankovitch spectrum.
(relative) greenhouse climate periods like the Triassic. Other Analyzing external controls on cyclic deposition, e.g., orbital
causal links between Milankovitch forcing and cyclic deposition forcing, requires an input data type that is able to directly reflect
on carbonate platforms must have existed. The global sensitivity the external signal. Orbital forcing in the Milankovitch band
to climate changes in the Milankovitch bandwidth during the triggers climate changes, which result in high-frequency, low-
Triassic can be assessed on the grounds of seasonal variations in amplitude sea-level changes. In combination with total subsid-
temperature, precipitation, atmospheric circulation, and oceanic ence, fluctuating sea level produces changes in accommodation
heat flux. Seasonal variations in these parameters serve as a proxy space. Temporal variations in accommodation space cannot be
for climate changes during Milankovitch cycles in one hemi- derived from lithofacies analysis, because lithofacies essentially
sphere. reflects paleobathymetry and not accommodation. In contrast to
Given the absence of polar ice sheets in the Triassic, Hay and accommodation, paleobathymetry is additionally influenced by
Leslie (1990) propose that changes in the volume of water stored localized, short-term changes in sediment accumulation, trig-
in groundwater reservoirs may have triggered eustatic sea-level gered for instance by biological, hydrodynamic, or paleoecologic
changes. Volume of stored water may have been controlled by variations. Subtidal changes in paleobathymetry may occur,
plate-tectonically driven changes in continental elevation (Trias- although the rate of accommodation has remained constant. On
sic: approx. 1.2–1.4 km, Hay et al. 1981). However, this process the large scale, Posamentier et al. (1988) and Schlager (1993) have
refers rather to sea-level changes of ≥ 1 My (third-order to second- demonstrated the fundamental difference between changes in
order sea-level changes). Donovan and Jones (1979) and Elder et paleobathymetry and accommodation. Bond and Kominz (1991)
al. (1994) suggested that, during high insolation, surface waters have described and analyzed the problem of using vertical facies
thermally expanded and contributed to sea-level changes. Schulz changes to infer accommodation and eustatic sea-level histories.
and Schäfer-Neth (1997) further developed this thermal-expan- They stressed that because of the effects of localized sediment
sion model for Mesozoic greenhouse climates. On the basis of accumulation rates on facies patterns, the interpretation of ac-
atmospheric general circulation models for the Triassic, they commodation, i.e., external controls on deposition, from vertical
proposed that warm, saline deep waters formed in summer when facies is far from straightforward. In order to correctly identify an
net evaporation rates were high. In winter, saline deep waters did orbitally forced sea-level signal, accommodation change and not
not form in the same hemisphere. However, warm, saline deep water-depth changes have to be analyzed.
water was formed throughout the year alternating between the Data on lithofacies rank or water-depth introduce a high-
two hemispheres. Schulz and Schäfer-Neth (1997) transferred frequency signal into the series, which is internally controlled by
this summer–winter situation to the hot–cold orbit situation sediment accumulation. The rank series of Preto et al. (2001)
during a Milankovitch cycle. Temperature variations of 2–4°C mixes external controls and internal controls on cyclic deposition.
between hot and cold orbits produced sea-level changes with an Internal control includes vertical successions of open and re-
amplitude of 1.7 m. Sea-level changes of this amplitude may well stricted subtidal facies (ranks 3, 4) and of supratidal facies (ranks
have controlled or contributed to the thinning-upward microcycles 1, 2). Of the 471 boundary conditions (rank boundaries) included
and higher-order cycles in the Latemar platform interior. Ther- in the series of Preto et al. (2001), only 39% (182 shallow subtidal
mal expansion of the water column because of non-orbitally to supratidal basic cycles) unequivocally indicate a change in
forced, high-frequency climatic changes also represents a poten- accommodation space. The remaining 61% (289) introduce a
tial driving force for sub-Milankovitch sea-level fluctuations in wide and laterally variable range of high-frequency signals,
the Triassic, resulting in the development of shallowing-upward which are probably due to internal control, to the series.
microcycles on carbonate-platform interiors. The individual thicknesses of accommodation cycles consti-
tute the exclusive data input to the spectral analyses presented in
Spectral Analysis this study. In contrast to Goldhammer et al. (1993), measured
sections do not include a model-dependent definition of
Data Types and Input.— microcycles. Where microcycles have been condensed beyond
recognition, no microcycles were defined. In such cases, it is
Spectral analysis studies in the Latemar platform interior preferable to miss beats rather than schematically introduce a
have used three different data types. Hinnov and Goldhammer model-specific stacking pattern like 1:5 into the series. Intervals in
INTEGRATED CYCLOSTRATIGRAPHY—THE LATEMAR (MIDDLE TRIASSIC, ITALY) 207

the Latemartürme section where microcycles have been con- Tukey, multi-taper harmonic, and spectral analysis. Signifi-
densed beyond recognition may include another 25–35 microcycles cant average ratios show little variation.
in addition to the 619 (min.) to 701 (max.) cycles actually mea- • the results of spectral analysis do not change if spectral power
sured. Apart from strongly condensed intervals, thinning-up- (Blackman–Tukey, multi-taper spectral analysis), amplitude
ward macrocycles as defined in this study include between two or significance levels from F-tests (multi-taper harmonic analy-
to six microcycles. sis) are considered.
• the same peaks are present in the power and amplitude
Power and Amplitude Spectra.— spectra for all sliding windows (subsets). According to Stage
(1999), this indicates that the cycles are real.
The identification of prominent frequencies in Blackman– • the model resulting from spectral analyses is well compatible
Tukey power spectra represents a straightforward procedure with the high-resolution biochronostratigraphic framework
that is based exclusively on the (relative) spectral power in each constrained by isotopic ages from the Latemar cyclic series,
sliding window. Frequencies above a power level of 2.5–3.0 were interpolated ages in current chronostratigraphic charts, the
selected. complete set of biostratigraphic data for the Latemar, and
Potential perturbations of the primary Milankovitch spec- various biostratigraphic schemes.
trum in the Latemar cyclic series include (1) nonstationary peri-
ods of sub-Milankovitch cyclicity; (2) hierarchical sub-Milanko- CONCLUSIONS
vitch cyclicities; (3) variations in the rate of deposition; (4)
autocyclic input in the series. These potential perturbation are not Individual shallowing-upward, meter-scale microcycles in
specific to the Latemar series but can be expected to exist in most the Latemar clearly reflect allocyclic control by changes in sea
cyclic carbonate-platform series. level. Individual cycles show moderate lateral variations in
Similarly to recent sub-Milankovitch cyclicities (D–O cycles), thickness, internal facies, and diagenetic overprint. More than
Triassic sub-Milankovitch cycle periods were probably less sta- 90% of macrocycles can be physically traced all over the plat-
tionary than Triassic Milankovitch cycle periods. Moderately form. Within laterally persistent macrocycles, 80–95% of all
shifting sub-Milankovitch cyclicity could raise the level of noise microcycles equally show lateral physical persistence. Except
in the Latemar series. Similarly to the Heinrich events and super- for the basal LCF, no stratal terminations exist on the microcycle
imposed Bond cycles in recent icehouse climate conditions, Trias- to macrocycle scale. The small diameter of the platform and the
sic sub-Milankovitch cyclicities of different orders may have been high average accommodation rates favored a simultaneous
superimposed on each other. The significant to highly significant vertical aggradation of the platform top. This allocyclic model
frequencies and amplitudes with an average ratio of 1:3.7 to 1:3.8 for the Latemar platform should not automatically be trans-
microcycles in Blackman–Tukey and multi-taper spectral and ferred to other Triassic or Mesozoic platforms with similar
harmonic analysis (Figs. 11A–B, 12A) possibly represent higher- subtidal to peritidal cycles and hierarchical cycle stacking pat-
order sub-Milankovitch cyclicity (Fig. 12B, 13.8–16.0 ky). This terns (e.g., Upper Triassic Dachstein Fm.). Each platform re-
ratio occurs in the same sliding windows irrespective of the quires detailed cyclostratigraphic logging, lateral tracing of
method of spectral analysis applied. cycles, and biochronostratigraphic and spectral analyses in
Short-term and long-term variations in the sedimentation order to develop a reliable forcing model. The Latemar repre-
rate may change the spectrum by increasing the signal-to-noise sents a model platform because it offers a unique combination
ratio. Reverse modeling for the Latemar platform (Zühlke et al., of features: well developed cyclicity, excellent outcrops for the
2003) indicates that the decompacted maximum sedimentation lateral physical tracing of cycles, abundant ammonoid faunas,
rates varied between 230 mm/ky (MTF) and 470 mm/ky (LCF), and volcanic ash layers intercalated in the platform interior.
i.e., for a factor of 2 during a time interval of 2.95 My (LCF–UTF; Basic meter-scale shallowing-upward microcycles in the
cf. Fig. 7, column E3). Because the cyclic series has been subdi- Latemar platform reflect sub-Milankovitch control. High-reso-
vided into sliding windows, variations in the sedimentation lution biochronostratigraphic data indicate clearly that basic
rate are small within single windows (0.5–0.9 My) and do not microcycles cannot be reconciled with precession forcing. The
increase the signal-to-noise ratio significantly. It is probable that biochronostratigraphic framework includes a large set of bio-
short-term sedimentation rates varied over 4.2 ky (single stratigraphic data from different studies, isotopic ages from the
microcycle) and 18.1–21.5 ky (1:4.3 to 1:5.1 bundled thinning- Latemar, and general data from current biochronostratigraphic
upward macrocycles). Sedimentation rates were high in the charts (Hardenbol et al., 1998). The period of each microcycle
subtidal part of a microcycle (LFT 1, after the initial lag time) ranges between ∆tav = 2.68 ky and ∆tmax ≤ 5.85 ky, compared to
compared to the supratidal top (LFTs 3–5). High average sedi- the Triassic short-precession period of ∆t = 17.78 ky. Spectral
mentation rates prevailed in the basal microcycles and reduced analyses show that the most probable microcycle period in the
sedimentation rates in the upper microcycles of thinning-up- Latemar platform interior is ∆t = 4.2 ky, which is less than a
ward macrocycles. Schwarzacher (1989) and Stage (1999) have quarter of the precessional period. These results are in clear
demonstrated that moderate random variations from the mean contrast to studies of, e.g., Goldhammer et al. (1993) and Preto
sedimentation rate alter the spectra slightly but original single et al. (2001), which assumed that individual microcycles or the
frequency peaks are still highly significant. Random small varia- shortest periodicity in rank series of the Latemar platform
tions in the sedimentation rate tend to attenuate high frequen- interior presented precession forcing of ∆t = 17.78 ky. These
cies especially. studies relied exclusively on the spectral analysis of series that
In summary, the spectral-analysis results of this study reflect covered a small part (one-third) of the cyclic succession pre-
external control. The following criteria indicate that the spectra served in the Latemar.
hold high confidence levels and that the selected frequencies and In the Latemar, cycle bundles of 1:4.3 to 1:5.1 shallowing-
amplitudes are real: upward microcycles do not reflect eccentricity forcing superim-
posed on precession forcing. Instead, the ratio reflects short and
• the same prominent frequencies and amplitudes occur with long precession forcing of 18 ky to 21 ky superimposed on sub-
all spectral-analysis methods applied, especially Blackman– Milankovitch forcing of 4.2 ky. At least for Mesozoic platform-
208 RAINER ZÜHLKE

interior successions, the interpretation of Milankovitch forcing REFERENCES


and equivalent time intervals from specific cycle stacking pat-
terns is unwarranted. Establishing orbital forcing in Mesozoic BECHSTÄDT, T., BRACK, P., PRETO, N., RIEBER, H., AND ZÜHLKE, R., 2003,
series requires a model-independent calibration of time by Fieldtrip to Latemar: Guidebook, Triassic Geochronology and Cyclo-
biochronostratigraphic data. If sufficiently high-resolution data stratigraphy—A Field Symposium, St. Christina/Val Gardena, Italy,
are not available, the time intervals implied by the orbital-forcing September 11–13, 72 p. (see also www.Latemar.sedimentology.info)
model can be compared to current biochronostratigraphic charts BERGER, A., AND LOUTRE, M.F., 1994, Astronomical forcing through geologi-
in order to check whether these time intervals are internally cal time, in de Boer, P.L., and Smith, D.G., eds., Orbital Forcing and
plausible. Cyclic Sequences: International Association of Sedimentologists, Spe-
Spectral analyses in this study have been both thickness- cial Publication 19, p. 15–24.
calibrated and time-calibrated, on the basis of high-resolution BLACKMAN, R.B., AND TUKEY, J.W., 1958, The Measurement of Power Spectra
biochronostratigraphic framework available for the Latemar. from the Point of View of Communication Engineering: New York,
Several spectral-analysis methods (Blackman–Tukey, multi-taper Dover Publications, 190 p.
spectral, and harmonic analysis) show highly similar significant BOND, G.C., HEINRICH, H., BROECKER, W., LABEYRIE, L., MCMANUS, J., ANDREWS,
ratios that are largely stationary over all seven subsets. The J., HUON, S., JANTSCHIK, R., CLASEN, S., SIMET, C., TEDESCO, K., KLAS, M.,
platform-interior cyclic series reveals higher-order significant BONANI, G., AND IVY, S., 1992, Evidence for massive discharges of
amplitudes at both ratios and time intervals indicative of orbital icebergs into the North Atlantic ocean during the last glacial period:
forcing in the Milankovitch band, when the microcycle period is Nature, v. 360, p. 245–249.
set to ∆t = 4.2 ky. This cyclostratigraphic, biochronostratigraphic, BOND, G.C., AND KOMINZ, M.A., 1991, Some comments on the problem of
and spectral-analysis model indicates that in addition to sub- using vertical facies changes to infer accommodation and eustatic sea-
Milankovitch forcing, short and long precession, short and long level histories with examples from Utah and the southern Canadian
obliquity, and short eccentricity forcing are present in the large- Rockies, in Franseen, E.K., Watney, W.L., Kendall, C.G.St.C., and
scale cycle bundlings. The Latemar cyclic series includes the Ross, W., eds., Sedimentary Modeling: Computer Simulations and
oldest explicit sub-Milankovitch signal and the oldest set of both Methods for Improved Parameter Definition: Kansas Geological Sur-
sub-Milankovitch and Milankovitch signals yet observed in the vey, Bulletin 233, p. 273–291.
geologic record. BOND, G.C., KOMINZ, M.A., AND BEVAN, J., 1991, Evidence for orbital forcing
The forcing of sub-Milankovitch cyclicity in the early Meso- of Middle Cambrian peritidal cycles, Wah Wah Range, south-central
zoic is yet unclear. Quaternary sub-Milankovitch cycles (e.g., Utah, in Franseen, E.K., Watney, W.L., Kendall, C.G.St.C., and Ross,
Dansgaard–Oeschger cycles) are linked to icehouse climate con- C.A., eds., Sedimentary Modeling: Computer Simulations and Meth-
ditions with polar ice caps and oceanic circulation patterns influ- ods for Improved Parameter Definition: Kansas Geological Survey,
enced by the current plate-tectonic configuration. In contrast, the Bulletin 233, p. 293–318.
early Mesozoic was characterized by (relative) greenhouse cli- BOND, G.C., KOMINZ, M.A., DEVLIN, W.J., AND BEAVAN, J., 1993, Evidence
mate conditions and the Pangea supercontinent plate-tectonic of astronomical forcing of the Earth´s climate in Cretaceous and
configuration. Therefore, Quaternary sub-Milankovitch cycles Cambrian times, in Anderson, D.M., and Eaton, G.P., eds., Geo-
cannot serve as direct analogies for (early) Mesozoic sub-Mi- logical Perspectives on Global Change: Tectonophysics, v. 222, p.
lankovitch cycles. In general, modern D–O cycles are caused by 295–315.
rapid changes in coupled ocean-atmosphere circulation patterns. BOSELLINI, A., AND ROSSI, D., 1974, Triassic carbonate buildups of the
Strong latitudinal temperature gradients and seasonal climate Dolomites, northern Italy, in Laporte, L.F., ed., Reefs in Time and
differences represent the major driver for modern oceanic–atmo- Space: Society of Economic Paleontologists and Mineralogists, Spe-
spheric circulation patterns and sub-Milankovitch cycles. Trias- cial Publication 18, p. 209–233.
sic climate was equally characterized by large latitudinal tem- BOSELLINI, A., CASTELLARIN, A., DOGLIONI, C., GUY, F., LUCCHINI, F., PERRI,
perature gradients and seasonal climate differences. Therefore, M.C., ROSSI, P.L., SIMBOLI, G., AND SOMMAVILLA, E., 1982, Magmatismo
the Triassic period held a very significant potential for fluctua- e tettonica nel Trias delle Dolomiti, in Castellarin, A., and Vai, G.B.,
tions in ocean–atmosphere circulation. eds., Guida alla Geologia del Sudalpino centro-orientale: Società
The most probable link between high-frequency, low-ampli- Geologica Italiana, Guide Geologiche Regionali, p. 189–210.
tude sea-level changes and both sub-Milankovitch and BOWRING, S.A., ERWIN, D.H., JIN, Y.G., MARTIN, M.W., DAVIDEK, K., AND
Milankovitch forcing in the Triassic is the thermal contraction WANG, W., 1998, Geochronology of the end-Permian mass extinction:
and expansion of the oceanic water column. Global seizure or Science, v. 280, p. 1039–1045.
release of water stored in groundwater reservoirs and associated BRACK, P., MUNDIL, R., OBERLI, F., MEIER, M., AND RIEBER, H., 1996, Biostrati-
changes in the runoff–evaporation–precipitation cycle may have graphic and radiometric age data question the Milankovitch charac-
triggered lower-frequency sea-level changes and contributed to teristics of the Latemar cycles (Southern Alps, Italy): Geology, v. 24,
higher-order cycles on carbonate platforms. p. 371–375.
BRACK, P., AND RIEBER, H., 1993, Towards a better definition of the Anisian/
ACKNOWLEDGMENTS Ladinian boundary: new biostratigraphic data and correlations of
boundary sections from the Southern Alps: Eclogae Geologicae
I am greatly indebted to Thilo Bechstädt, who initiated this Helvetiae, v. 86, p. 415–527.
project and supported it in many ways. Peter Brack, Hans Rieber, BRACK, P., AND RIEBER, H., 1994, The Anisian/Ladinian boundary: retro-
John Reijmer, Wolfgang Schwarzacher, and Alton Brown are spective and new constraints: Albertiana, v. 13, p. 25–36.
thanked for discussions and comments. Alfred Fischer and two BROECKER, W., AND DENTON, G.H., 1989, The role of ocean–atmosphere
anonymous reviewers provided many valuable comments. reorganizations in glacial cycles: Geochimica et Cosmochimica Acta,
Claudia Seibold detected the ammonoid fauna below the summit v. 53, p. 2465–2501.
of the Latemartürme. The author is very grateful to the Gabrielli BROECKER, W.S., BOND, G.C., KLAS, M., CLARK, E.A., AND MCMANUS, J., 1992,
family from the Rifugio Torre di Pisa on the Latemar for their Origin of the northern Atlantic’s Heinrich events, in Kelts, K.R., ed.,
hospitality and logistical support. Part of this study was funded Past and Present Climate Dynamics; Reconstruction of Rates of
by the German Research Fund (DFG, Be 641/28). Change: Climate Dynamics, v. 6, p. 265–273.
INTEGRATED CYCLOSTRATIGRAPHY—THE LATEMAR (MIDDLE TRIASSIC, ITALY) 209

CISNE, J.L., 1986, Earthquakes recorded stratigraphically on carbonate GOLDHAMMER, R.K., OSWALD, E.J., AND DUNN, P.A., 1994, High-frequency,
platforms: Nature, v. 323, p. 320–322. glacio-eustatic cyclicity in the Middle Pennsylvanian of the Paradox
DE ZANCHE, V., GIANOLLA, P., MANFRIN, S., MIETTO, P., AND ROGHI, G., 1995, Basin: an evaluation of Milankovitch forcing, in de Boer, P.L., and
A Middle Triassic back-stepping carbonate platform in the Dolomites Smith, D.G., eds., Orbital Forcing and Cyclic Sequences: International
(Italy): sequence stratigraphy and biochronostratigraphy: Memorie Association of Sedimentologists, Special Publication 19, p. 243–283.
di Scienze Geologiche, v. 47, p. 135–155. GRADSTEIN, F.M., AGTERBERG, F.P., OGG, J.G., HARDENBOL, J., VAN VEEN, P.,
DONOVAN, D.T., AND JONES, E.J.W., 1979, Causes of world-wide changes THIERRY, J., AND HUANG, Z., 1994, A Mesozoic time scale: Journal of
in sea-level: Geological Society of London, Journal, v. 136, p. 187– Geophysical Research, v. 99, p. 24,051–24,074.
192. GRADSTEIN, F.M., AGTERBERG, F.P., OGG, J.G., HARDENBOL, J., VAN VEEN, P.,
DRUMMOND, C.N., AND WILKINSON, B.H., 1993, Carbonate cycle stacking THIERRY, J., AND HUANG, Z., 1995, A Triassic, Jurassic and Cretaceous
patterns and hierarchies of orbitally forced eustatic sea level change: time scale, in Berggren, W.A., Kent, D.V., and Aubry, M.-P., eds.,
American Association of Petroleum Geologists, Bulletin, v. 63, p. Geochronology, Time Scales and Global Stratigraphic Correlation:
369–377. SEPM, Special Publication 54, p. 95–126.
DUNN, P.A., 1992, Cyclic stratigraphy and early diagenesis; an example GROTZINGER, J.P., 1986, Upward shallowing platform cycles: a response to
from the Triassic Latemar Platform, northern Italy: Unpublished 2.2 billion years of low-amplitude, high-frequency (Milankovitch
Ph.D. Thesis, John Hopkins University, Baltimore, 865 p. band) sea level oscillations: Paleoceanography, v. 1, p. 403–416.
EGENHOFF, S., PETERHÄNSEL, A., BECHSTÄDT, T., ZÜHLKE, R., AND GRÖTSCH, J., HARDENBOL, J., THIERRY, J., FARLEY, M.B., JACQUIN, T., DE GRACIANSKY, P.-C.,
1999, Facies architecture of an isolated carbonate platform: tracing AND VAIL, P.R., 1998, Mesozoic and Cenozoic sequence chronostrati-
the cycles of the Latemar (Middle Triassic, northern Italy): Sedimen- graphic framework of European basins, in de Graciansky, P.-C.,
tology, v. 46, p. 893–912. Hardenbol, J., Jaquin, T., and Vail, P.R., eds., Mesozoic and Cenozoic
ELDER, W.P., GUSTASON, E.R., AND SAGEMANN, B.B., 1994, Correlation of Sequence Stratigraphy of European Basins: SEPM, Special Publica-
basinal carbonate cycles to nearshore parasequences in the late tion 60, p. 3/15.
Cretaceous Greenhorn seaway, Western interior, USA: Geological HARDIE, L.A., 1986, Stratigraphic Models for carbonate tidal-flat deposi-
Society of America, Bulletin, v. 106, p. 892–902. tion, in Hardie, L.A., and Shinn, E.A., eds., Carbonate Depositional
ENOS, P., AND SAMANKASSOU, E., 1998, Lofer cyclothems revisited (Late Environments, Part 3: Tidal Flats: Colorado School of Mines, Quar-
Triassic, Northern Alps, Austria): Facies, v. 38, p. 207–228. terly, v. 81, p. 59–73.
FISCHER, A.G., 1964, The Lofer cyclothems of the Alpine Triassic: Kansas HARDIE, L.A., AND SHINN, E.A., 1986, Carbonate depositional environ-
Geological Survey, Bulletin 169, p. 107–149. ments: Colorado School of Mines, Quarterly, v. 81, 74 p.
FISCHER, A.G., 1986, Climatic rhythms recorded in strata: Annual Review HARDIE, L.A., BOSELLINI, A., AND GOLDHAMMER, R.K., 1986, Repeated sub-
of Earth and Planetary Sciences, v. 14, p. 351–376. aerial exposure of subtidal carbonate platforms, Triassic, northern
GAETANI, M., FOIS, E., JADOUL, F., AND NICORA, A., 1981, Nature and Italy: evidence for high-frequency sea-level oscillations on a 104 year
evolution of the middle Triassic carbonate buildups in the Dolo- scale: Paleoceanography, v. 1, p. 447–457.
mites (Italy): Marine Geology, v. 44, p. 25–57. HARDIE, L.A., WILSON, E.M., AND GOLDHAMMER, R.K., 1991, Cyclostratigra-
GIANOLLA, P., DE ZANCHE, V., AND MIETTO, P., 1998, Triassic sequence stra- phy and dolomitization of the Middle Triassic Latemar buildup, the
tigraphy in the Southern Alps (northern Italy): definition of sequences Dolomites, northern Italy: Dolomieu Conference on Carbonate Plat-
and basin evolution, in de Graciansky, P.C., Hardenbol, J., Jacquin, T., forms and Dolomitization, September, St. Ullrich/Gröden, Guide-
and Vail, P.R., eds., Mesozoic and Cenozoic Sequence Stratigraphy of book Excursion F, p. 1–56.
European Basins: SEPM, Special Publication 60, p. 719–747. HARLAND, W.B., ARMSTRONG, R.L., COX, A.V., CRAIG, L.E., SMITH, A.G., AND
GINSBURG, R.N., 1971, Landward movement of carbonate mud: a new SMITH, D.G., 1990, A Geologic Time Scale 1989: Cambridge, U.K.,
model for regressive cycles in carbonates (abstract): American Asso- Cambridge University Press, 263 p.
ciation of Petroleum Geologists, Bulletin, v. 55, p. 340. HARLAND, W.B., COX, A.V., LLEWELLYN, P., THICKTON, C.A., SMITH, A.G., AND
GOLDHAMMER, R.K., 1987, Platform carbonate cycles, middle Triassic of WALTERS, R., 1982, A Geological Time Scale: Cambridge, U.K., Cam-
northern Italy: the interplay of local tectonics and global eustasy: bridge University Press, 131 p.
Unpublished Ph.D. Thesis, John Hopkins University, Baltimore, HARRIS, M.T., 1994, The foreslope and toe-of-slope facies of the middle
468 p. Triassic Latemar buildup (Dolomites, northern Italy): Journal of
GOLDHAMMER, R.K., DUNN, P.A., AND HARDIE, L.A., 1987, High-frequency Sedimentary Research, v. B64, p. 132–145.
glacio-eustatic oscillations with Milankovitch characteristics re- HARRIS, M.T., KERANS, C., AND BEBOUT, D.G., 1993, Ancient outcrop and
corded in northern Italy: American Journal of Science, v. 287, p. 853– modern examples of platform carbonate cycles—implications for
892. understanding subsurface correlation and understanding reservoir
GOLDHAMMER, R.K., DUNN, P.A., AND HARDIE, L.A., 1990, Depositional heterogeneity, in Loucks, R.G., and Sarg, J.F., eds., Carbonate Se-
cycles, composite sea-level changes, cycle stacking patterns, and the quence Stratigraphy: Recent Developments and Applications: Ameri-
hierarchy of stratigraphic forcing: Geological Society of America, can Association of Petroleum Geologists, Memoir 57, p. 475–492.
Bulletin, v. 102, p. 535–562. HAY, W.W., BARRON, E.J., SLOAN, J.L., AND SOUTHAM, J., 1981, Continental
GOLDHAMMER, R.K., AND HARRIS, M.T., 1989, Eustatic controls on the drift and the global pattern of sedimentation: Geologische Rundschau,
stratigraphy and geometry of the Latemar buildup (Middle Trias- v. 70, p. 302–315.
sic), the Dolomites of northern Italy, in Crevello, P.D., Wilson, J.L., HAY, W.W., AND LESLIE, M.A., 1990, Could possible changes in global
Sarg, J.F., and Read, J.F., eds., Controls on Carbonate Platform and groundwater reservoir cause eustatic sea-level fluctuations?, in Revelle,
Basin Development: SEPM, Special Publication 44, p. 323–338. R., ed., Sea-Level Change: National Research Council, Studies in
GOLDHAMMER, R.K., HARRIS, M.T., DUNN, P.A., AND HARDIE, L.A., 1993, Geophysics, p. 161–170.
Sequence stratigraphy and system tract development of the Latemar HEINRICH, H., 1988, Origin and consequences of cyclic ice rafting in the
Platform, middle Triassic of the Dolomites (Northern Italy): outcrop Northeast Atlantic Ocean during the past 130,000 years: Quaternary
calibration keyed by cycle stacking patterns, in Loucks, R.G., and Research, v. 29, p. 142–152.
Sarg, J.F., eds., Carbonate Sequence Stratigraphy: Recent Develop- HILGEN, F.J., KRIJGSMAN, W., LANGEREIS, C.G., AND LOURENS, L.J., 1997,
ments and Applications: American Association of Petroleum Geolo- Breakthrough made in dating of the geological record: Eos, Transac-
gists, Memoir 57, p. 353–388. tions, American Geophysical Union, v. 78, p. 285–289.
210 RAINER ZÜHLKE

HINNOV, L.A., AND GOLDHAMMER, R.K., 1991, Spectral analysis of the Middle POSAMENTIER, W., JERVEY, M.T., AND VAIL, P.R., 1988, Eustatic controls on
Triassic Latemar Limestone: Journal of Sedimentary Petrology, v. 61, clastic deposition I—conceptual framework, in Wilgus, C.K., Hastings,
p. 1173–1193. B.S., Kendall, C.G.St.C., Posamentier, H.W., Ross, C.A., and Van
HÜSSNER, H.M., ROESSLER, J., BETZLER, R., AND PETSCHICK, R., 2000, Milanko- Wagoner, J.C., eds., Sea-Level Changes: An Integrated Approach:
vitch-Steuerung ohne orbitale Steuerung?: Gesellschaft der Geologie- SEPM, Special Publication 42, p. 109–124.
und Bergbaustudenten Österreichs, Mitteilungen, v. 43, p. 66. PRATT, B.R., JAMES, N.P., AND COWAN, C.A., 1992, Peritidal carbonates, in
KOERSCHNER, W.F., AND READ, J.F., 1989, Field and modelling studies of Walker, R.G., and James, N.P., eds., Facies Models: Response to Sea
Cambrian carbonate cycles, Virginia Appalachians: Journal of Sedi- Level Change, Geological Association of Canada, p. 303–322.
mentary Petrology, v. 59, p. 654–687. PRETO, N., AND HINNOV, L., 2000, Semi-automated correlation and the
MASETTI, D., AND TROMBETTA, G.L., 1998, L´eredita anisica nella nascita ed cyclicity of the Latemar platform (Middle Triassic, Southern Alps)
evoluzione delle piattaforme medio-triassiche delle dolomiti (abstract): International Association of Sedimentologists, 20th Regional
occidentali: Memorie di Scienze Geologiche, v. 50, p. 213–237. Meeting, 13–15 September, Dublin, Ireland, Abstract Volume, p. 127.
MIALL, A.D., 1997, The Geology of Stratigraphic Sequences: Berlin, Springer, PRETO, N., MIETTO, P., AND MANFRIN, S., 2002, Ammonoid biostratigraphy
433 p. of the Latemar platform and its significance for the A/L boundary
MIETTO, P., AND MANFRIN, S., 1995, A high-resolution ammonoid standard (abstract), in Piros, O., ed., I.U.G.S. Subcommission on Triassic stratig-
scale in the Tethys realm. A preliminary report: Société Géologique de raphy, STS/IGCP 467 field meeting: Hungarian Geological Society,
France, Bulletin, v. 166, p. 539–563. 5–8 September, Veszprem, Hungary, Abstract Volume, p. 21.
MILANKOVITCH, M., 1941, Kanon der Erdbestrahlung und seine Anwendung PRETO, N., HINNOV, L.A., HARDIE, L.A., AND DE ZANCHE, V., 2001, Middle
auf das Eiszeitenproblem: Royal Serbian Academy, Section of Math- Triassic orbital signature recorded in the shallow marine Latemar
ematics and Natural Sciences, Special Publication, v. 132. carbonate buildup (Dolomites, Italy): Geology, v. 29, p. 1123–1126.
MONTAÑEZ, I.P., AND OSLEGER, D.A., 1993, Parasequence stacking patterns, RASMUSSEN, T.L., THOMSEN, E., AND VAN WEERING, T.C.E., 1998, Cyclic sedi-
third-order accommodation events, and sequence stratigraphy of mentation on the Faeroe rift 53-10 ka related to climatic variations, in
Middle to Upper Cambrian platform carbonates, Bonanza King For- Stroker, M.S., Eavens, D., and Cramp, A., eds., Geological Processes on
mation, Southern Great Basin, in Loucks, R.G., and Sarg, J.F., eds., Continental Margins: Sedimentation, Mass-Wasting and Stability:
Carbonate Sequence Stratigraphy: Recent Developments and Appli- Geological Society of London, Special Publication 129, p. 225–267.
cations: American Association of Petroleum Geologists, Memoir 57, READ, J.F., 1989, Controls on evolution of Cambrian–Ordovician passive
p. 305–326. margin, U.S. Appalachians, in Crevello, P.D., Sarg, J.F., Read, J.F., and
MONTAÑEZ, I.P., AND READ, J.F., 1992, Eustatic control on early dolomitiza- Wilson, J.L., eds., Controls on Carbonate Platform and Basin Develop-
tion of cyclic peritidal carbonates: evidence from the early Ordovician ment: SEPM, Special Publication 44, p. 147–166.
Upper Knox Group, Appalachians: Geological Society of America, READ, J.F., AND GOLDHAMMER, R.K., 1988, Use of Fischer plots to define
Bulletin, v. 104, p. 872–886. third-order sea-level curves in Ordovician peritidal cyclic carbonates,
MUNDIL, R., BRACK, P., AND LAURENZI, M.A., 1996a, High-resolution U/Pb Appalachians: Geology, v. 16, p. 895–899.
single-zircon age determinations: new constraints on the timing of READ, J.F., KERANS, C., WEBER, L.J., SARG, J.F., AND WRIGHT, F.M., 1995,
Middle Triassic magamatism in the Southern Alps (abstract): Società Milankovitch sea-level changes, cycles, and reservoirs on carbonate
Geologica Italiana, 78a Riunione Estiva, September 16–18, San platforms in greenhouse and icehouse worlds: SEPM, Short Course
Cassiano, Italy, Abstract Volume. Notes 35, 203 p.
MUNDIL, R., BRACK, P., MEIER, M., RIEBER, H., AND OBERLI, F., 1996b, High RÜFFER, T., AND ZÜHLKE, R., 1995, Sequence stratigraphy and sea-level
resolution U–Pb dating of middle Triassic volcaniclastics: time-scale changes in the Early to Middle Triassic of the Alps: a global compari-
calibration and verification of tuning parameters for carbonate sedi- son, in Haq, B.U., ed., Sequence Stratigraphy and Depositional Re-
mentation: Earth and Planetary Science Letters, v. 141, p. 137–151. sponse to Eustatic, Tectonic and Climatic Forcing: Dordrecht, The
MUNDIL, R., ZÜHLKE, R., BECHSTADT, T., BRACK, P., EGENHOFF, S., MEIER, M., Netherlands, Kluwer Academic Publishers, p. 161–207.
OBERLI, F., PETERHÄNSEL, A., AND RIEBER, H., 2003, Cyclicities in Triassic SADLER, P.M., OSLEGER, D.A., AND MONTAÑEZ, I.P., 1993, On the labeling,
platform carbonates: synchronizing radio-isotopic and orbital clock: length and objective basis of Fisher plots: Journal of Sedimentary
Terra Nova, in press. Petrology, v. 63, p. 360–368.
OSLEGER, D.A., AND READ, J.F., 1991, Relation of eustasy to stacking patterns SANDER , B., 1936, Beiträge zur Kenntnis der Anlagerungsgefüge
of meter-scale carbonate cycles, late Cambrian, U.S.A: Journal of (Rhythmische Kalke und Dolomite aus der Trias); II, Südalpine
Sedimentary Petrology, v. 61, p. 1225–1252. Beispiele, Hauptdolomit, Allgemeines: Mineralogische und
PAILLARD, D., LABEYRIE, L., AND YIOU, P., 1996, Macintosh program performs Petrographische Mitteilungen, v. 48, p. 141–209.
time-series analysis: Eos, Transactions, American Geophysical Union, SATTERLY, A.K., 1996, Cyclic sedimentation in the upper Triassic Dachstein
v. 77, p. 379. Limestone, Austria: the role of patterns of sediment supply and
PALFY, J., PARRISH, R.R., AND VÖRÖS, A., 2002, Integrated U–Pb geochronol- tectonics in a platform–reef–basin system: Journal of Sedimentary
ogy and ammonoid biochronology from the Anisian/Ladinian GSSP Research, v. 66, p. 307–323.
candidate section at Felsöörs (abstract), in Piros, O., ed., I.U.G.S SCHLAGER, W., 1993, Accommodation and supply—a dual control on
Subcommission on Triassic stratigraphy, STS/IGCP 467 field meet- stratigraphic sequences: Sedimentary Geology, v. 86, p. 111–136.
ing: Hungarian Geological Society, 5–8 September, Veszprem, Hun- SCHULZ, M., AND SCHÄFER-NETH, C., 1997, Translating Milankovitch climate
gary, Abstract Volume, p. 28. forcing into eustatic fluctuations via deep water expansion: a concep-
PALMER, A.R., 1983, The Decade of North American Geology, 1983 geo- tual link: Terra Nova, v. 9, p. 228–231.
logic time scale: Geology, v. 11, p. 503–504. S CHWARZACHER , W., 1947, Über die sedimentäre Rhythmik des
PEPER, T., AND CLOETINGH, S., 1995, Autocyclic perturbations of orbitally Dachsteinkalks von Lofer: Vienna, Geologische Bundes-Anstalt,
forced signals in the sedimentary record: Geology, v. 23, p. 937– Verhandlungen, v. H10-12, p. 175–188.
940. SCHWARZACHER, W., 1989, Milankovitch cycles and the measurement of
PETERHÄNSEL, A., AND EGENHOFF, S., 2000, Cyclomania—Milankovitch cycles time: Terra Nova, v. 15, p. 405–408.
in the Triassic of northern Italy refuted (abstract): International Asso- SCHWARZACHER, W., AND HAAS, J., 1986, Comparative statistical analysis of
ciation of Sedimentologists, 20th Regional Meeting, September 13–15, some Hungarian and Austrian Upper Triassic peritidal carbonate
Dublin, Ireland, Abstract Volume, p. 46. sequences: Acta Geologica Hungarica, v. 29, p. 175–196.
INTEGRATED CYCLOSTRATIGRAPHY—THE LATEMAR (MIDDLE TRIASSIC, ITALY) 211

SLOAN, R.J., AND WILLIAMS, B.P.J., 1991, Volcano-tectonic control of offshore


to tidal-flat regressive cycles from the Dunquin Group (Silurian) of
southwest Ireland, in Macdonald, D.I.M., ed., Sedimentation, Tecton-
ics and Eustasy: Sea-Level Changes at Active Margins: International
Association of Sedimentologists, Special Publication 12, p. 105–119.
STAGE, M., 1999, Signal analysis of cyclicity in Maastrichtian pelagic chalks
from the Danish North Sea: Earth and Planetary Science Letters, v.
173, p. 75–90.
STEIGER, R., BICKEL, R.A., AND MEIER, M., 1993, Conventional U–Pb dating
of single fragments of zircon for petrogenetic studies of Phanerozoic
granitoids: Earth and Planetary Science Letters, v. 115, p. 197–209.
STRASSER, A., 1988, Shallowing upward sequences in Purbeckian peritidal
carbonates (lower Cretaceous, Swiss and French Jura Mountains):
Sedimentology, v. 35, p. 369-383.
THOMSON, D.J., 1982, Spectrum estimation and harmonic analysis: IEEE
Proceedings, v. 70, p. 1055-1096.
VÖRÖS, A., 1998, Triassic ammonoids and biostratigraphy of the Balaton
Highland: Studia Naturalis Magyar Természettudományi Múzeum,
v. 12.
WILSON, K.M., POLLARD, D., HAY, W.W., THOMPSON, S.L., AND WOLD, C.N.,
1994, General circulation model simulations of Triassic climates;
preliminary results, in Klein, G.D., ed., Pangea—Paleoclimate, Tec-
tonics, and Sedimentation during Accretion, Zenith, and Breakup of
a Supercontinent: Geological Society of America, Special Paper 288, p.
91–116.
WOLFRAM, S., 1996, The Mathematica Book: Wolfram Media/Cambridge
University Press, 1403 p.
YANG, H., AND KOMINZ, M.A., 1999, Testing periodicity of depositional
cyclicity, Cisco Group (Virgilian and Wolfcampian), Texas: Journal of
Sedimentary Research, v. 69, p. 1209–1231.
YIOU, P., GHIL, M., JOUZEL, J., PAILLARD, D., AND VAUTARD, R., 1994, Nonlinear
variability of the climatic system, from singular and power spectra of
Late Quaternary records: Climate Dynamics, v. 9, p. 371–389.
YU, H., 1995, Mathematica Time Series Pack—Reference and User´s
Guide: Champaign, Illinois, Wolfram Research Inc., 168 p.
ZÜHLKE, R., 2000, Fazies, hochauflösende Sequenzstratigraphie und
Beckenentwicklung im Anis (mittlere Trias) der Dolomiten (Südalpin,
N-Italien): Gaea Heidelbergensis, v. 6, p. 1–368.
ZÜHLKE, R., AND BECHSTÄDT, T., 2000, High-resolution cyclostratigraphy
and architecture of a carbonate platform interior—the Latemar, Middle
Triassic, Southern Alps (abstract): International Association of Sedi-
mentologists, 20th Regional Meeting, 13–15 September, Dublin, Ab-
stract Volume, p. 68.
ZÜHLKE, R., BECHSTÄDT, T., AND MUNDIL, R., 2001, Sub-Milankovitch and
Milankovitch cyclicity in an model carbonate platform—Latemar,
Triassic, Italy (abstract): American Association of Petroleum Geolo-
gists, Annual Meeting, June 3–6, Denver, Abstract Volume, p. A226.
ZÜHLKE, R., BECHSTADT, T., AND MUNDIL R., 2003, Sub-Milankovitch and
Milankovitch forcing on a model Mesozoic carbonate platform—the
Latemar (Middle Triassic, Italy): Terra Nova, in press.
ZÜHLKE, R., BECHSTÄDT, T., BRACK, P., MUNDIL, R., AND RIEBER, H., 2000, Die
Latemar-Kontroverse: neue Daten zur Geometrie, zeitlichen
Entwicklung und Interpretation lagunärer Zyklen: Gesellschaft der
Geologie- und Bergbaustudenten in Österreich, Mitteilungen, v. 43,
p. 154.
LATEMAR CYCLO-, BIO-, AND CHRONOSTRATIGRAPHY
RAINER ZÜHLKE

INTEGRATED CYCLOSTRATIGRAPHY OF A MODEL MESOZOIC CARBONATE PLATFORM–


THE LATEMAR (MIDDLE TRIASSIC, ITALY)
In D’Argenio, B., Fischer, A., Premoli Silva, I., Weissert, H., and Ferreri, V., eds., 2004,
Cyclostratigraphy: Approaches and Case Histories, SEPM Special Publication 81.
Chart 1

Section from Section from Section Section Section Section Section Bio-/Chronostratigr.
GOLDHAMMER (1987) PRETO ET AL. (2001)
Valsordaspitze Erzlahnspitze Latemartürme Valle de la Vècia Spiz de la Cagnota
Cima Valsorda Cima Forcellone Torri di Latemar Brack and Rieber (1993)
S- and NW-Latemar northern Latemar northeastern Latemar eastern Latemar Brack et al. (1996)
southwestern Latemar northwestern Latemar northern Latemar LCF to UCF LPF to MTF Mundil et al. (1996b)
LCF to UCF MTF to UCF
LTF to MTF MTF to UCF LTF to UTF 245 m 317 m Zühlke et al. (2000, 2003)
322 m 160 m
229 m 194 m 469 m Mundil et al. (2003)

erosional surface
at the summit of

A1 A2 A3 A4 A5 the Latemartürme
2842 m a.sl. A6 A7
base
wind7

460.00

lower Curionii Zone

“Ladinian”
early “Fassanian”
Upper Tepee Facies ( UTF )
450.00

Upper Tepee Facies ( UTF )


440.00

UTF 430.00
For Comparison Upper Tepee
Section according to PRETO et
al. (2001), Geological Society of
America; Data Repository item
Facies
2001129; p. 2-3; at www.geoso
ciety.org/ pubs/ft2001.htm

a) shallowing-, deepening-upward cycles


420.00

PRETO et al. (2001): Middle


Triassic orbital signature
recorded in the shallow marine
Latemar carbonate buildup
(Dolomites, Italy). - Geology, 29,
12, p. 1123-1126

rank 1: caliches, soils top


410.00 wind6
rank 2: supratidal flat
rank 3: restricted, subtidal lagoon
rank 4: open, subtidal lagoon
Shallowing and deepening-
upward cycles as well as cycle
numbers have not been in-
dicated by PRETO et al., but were
defined by the author based on 400.00
the values given in their data
repository (p. 2-3)

For Comparison The section by PRETO et al. is corre-


lated to the cyclostratigraphic units of
this study based on the MTF/UCF
boundary. Preto et al. indicate, that 390.00
Section according to b) alternative cyclostratigraphy (based on thin. upward trends) the section covers the uppermost MTF
a) cyclostratigraphy according to GOLDHAMMER (1987)

GOLDHAMMER (1987), to basal UTF of Egenhoff et al. (2001)


Appendix #4 and Ap- (Fig.1) and that the base of the section
pendix #2; (cp. GOLD- is ca. 10 m below the LCF/TF (sic,
HAMMER et al. 1989, MTF/UCF) boundary sensu Egenhoff
1991, 1993; HINNOV & et al. 1999) (data repository, p. 2). In
GOLDHAMMER 1991). fact, the section finishes in the upper
UCF.
GOLDHAMMER (1987): 380.00
Platform carbonate cy-
cles, Middle Triassic of
Northern Italy: the inter- Rank
play of local tectonics 4 3 2 1
and global eustasy. -
468 p., Ph.D. disserta- 158.12
“Upper Teepe Facies” (sic)

tion, The John Hopkins 156.35


University/Baltimore. 155.47 370.00
153.94
L-13 152.93
151.94
L-FSI 151.18
150.39
150.00
Correlation of sepa- 148.82
rately measured sec- 147.81
tions by Goldhammer 147.08
146.16
145.54 360.00
144.54
143.85
The section by Goldhammer is corre- 142.38
lated according to major cyclostrati-
graphic boundaries (LCF, TF, UCF) to 140.32 base
140.00 wind7
the sections presented in this study. top
138.44
The definition of boundaries by Gold- 137.76 wind5
hammer is based on the occurence of 136.89
136.19
teepes. It is not equivalent to the def- 350.00
inition of boundaries in this study. 134.55
133.65
132.88
132.04
erosional surface
at the summit of 130.00
the Erzlahnspitze 128.70
+130 m UCF 127.68
("inaccessible")
L-15 at the
("reconnais- Latemartürme 340.00
sance only") 124.56
track
511
121.31
120.56
317.40 120.00

116.36
115.22 330.00
114.27
311.10
112.44

Upper Cyclic Facies ( UCF )


308.30 110.80

early middle Triassic / “Anisian”


110.00
108.61

Upper Cyclic Facies ( UCF )


erosional surface
at the summit of
105.41 the Erzlahnspitze 320.00

Secedensis Zone / late “Illyrian”


104.75 2750 m a.sl.
301.60 104.05
103.25

294.80
100.00

97.07

95.17
94.47
93.59
UCF 190.00

310.00

90.49 Upper Cyclic


Upper Cyclic Facies ( UCF ) sensu Goldhammer

Upper Cyclic Facies ( UCF )

90.00
180.00
88.46

283.10
86.35
85.71
Facies 300.00
reef to slope facies
below Col Cornon
2625 m a.sl.
84.66
base
wind6

no cyclicity - m-scale to massive bedding

no cyclicity - m-scale to massive bedding


cycle stacking pattern vanishes due to oblique
top

angle approach to vertical platform margin


wind4
80.70
80.00 170.00 240.00
275.90 77.86
76.86
76.19
290.00 LAT-32
73.29
32 241.5 +- 0.4
0.6 Ma
L AT-
72.37

LAT-32
71.42
Upper Cyclic Facies ( UCF )

70.70 r
Laye U-Pb isotopic age from single
70.00 an. 230.00
160.00
V olcan. Layer V olc zircons in volcaniclastic layer
Mundil et al. (2003)
265.10 67.27
Median age based on normal
65.72 280.00 distribution of U concentrations
64.01 and/or Th/U rations:
241.5 +0.4/-0.6 Ma
259.60 Including additional ages de-
60.75 R viating from normal distribution:
R 241.7 +1.5/-0.7 Ma
60.00 220.00
150.00
57.88 Valsordaspitze

Upper Cyclic Facies ( UCF )


255.60
volcaniclastic layer above
56.43
253.60 55.58 base of section
270.00
54.48 Erzlahnspitze, sampled
53.07 Level 160.20 m
249.90 Latemartürme, projected
51.45
along traced physical surface
50.00 210.00 Level 284.60 m
140.00
48.48 Valle della Veccia
Spiz de la Cagnota
45.64 volcaniclastic layer above
243.20 260.00
44.81 base of section
43.77
42.94

40.78
40.00 200.00
130.00
235.90 38.34
37.64
36.56
35.06 250.00
231.50 34.01
32.82
31.62
227.90 30.00 120.00 190.00
28.76

26.95 base
224.20 wind5
25.14 N 240.00 N
222.60
N
23.19 top
219.70 21.99 wind3
20.94 Statistics
217.80 20.00
(LCF to UTF)
110.00 180.00
18.55

a) max. number of micro- and macrocycles

b) min. number of micro- and macrocycles


215.00 17.65
16.83 a) maximum number of
cycles
15.25 230.00
- microcycles: 699
14.09 - macrocycles: 156
210.60 12.96
erosional surface
209.20 11.75 at the summit of b) minimum number of
10.67 the Valsordaspitze cycles
207.60 10.00 2752 m a.sl.
170.00
100.00 M M M - microcycles: 621
L-13 8.17 - macrocycles: 135
L-FSI 6.43 c) average thickness
5.42 220.00
MTF

4.54 - microcycle (min.-


201.30 max.): 0.66-0.74 m
3.17
2.66 - macrocycle (min.-
220.00 max.): 2.95-3.41 m
1.66
198.50
0.00
90.00 160.00 (including projections
from reference section
MTP
Latemartürme)
Middle Tepee Facies ( MTF )

-0
210.00
192.10 Statistics
191.00 (LCF to UTF) sparse outcrops
190.10 for 45 m
a) shallowing-, deepening-upward cycles

210.00 followed by
189.30
188.50 reef and
a) maximum number of cycles slope facies
187.20 80.00 150.00
- total number of microcy-
cles: 215
184.70 - shallowing-upward micro-
J J J
cycles: 182 200.00
181.80 - deepening-upward micro-
cycles: 33 J J
310.00
200.00
179.10 b) minimum number of cycles
178.00 - total number of microcy-
177.10 cles: 182, 33 of which in- 70.00 140.00

175.40 clude a deepening upward


trend in their basal part.
173.60
Cycle numbers calculated 190.00
171.60 from Preto et al. (2001, data
Middle Tepee Facies ( MTF )

170.50 repository, p. 2-3) by the au- 300.00


169.80 thor. 190.00
168.50
base
167.30 60.00 wind4 130.00
166.10
Tepee Facies ( TF ) sensu Goldhammer

top
wind2

180.00 Middle Tepee Facies ( MTF )

early middle Triassic / “Anisian”


290.00
Middle Tepee Facies ( TF )

180.00
MTP LAT-30 LAT- LAT-30
-1
50.00 30 LAT-3 120.00
Volcaniclastic Layer V olc
an. L
ayer 0 LAT-30 241.2 +0.7
- 0.6 Ma
Secedensis Zone / late “Illyrian”
V olca track
n. La 518
yer
Volcan. Layer U-Pb isotopic age from single
170.00
zircons in volcaniclastic layer
151.40 Mundil et al. (2003)
280.00
170.00 Valsordaspitze, projected
149.20
Middle Tepee Facies ( MTF )

Level 178.90 m (?)


110.00
Middle Tepee Facies ( MTF )

40.00 Erzlahnspitze, projected


145.90 Level 51.00 m
145.00
Latemartürme, sampled
143.20 160.00 Level 173.00 m
141.60 Valle della Veccia, projected
270.00 Level 117.40 m
139.50 160.00
Spiz de la Cagnota, projected
Level 284.50 m
137.40 30.00 100.00

150.00
MTP
-2 260.00
150.00
128.90

20.00 90.00
126.70
Middle Tepee Facies ( MTF )

125.10
123.90
MTP 140.00
122.10 -3
121.20
250.00
L-FSII 140.00

L-10
10.00 80.00
116.60
115.10
113.70
130.00

111.20 base
wind3 240.00
109.60 130.00
top
wind1
0.00 70.00

105.20

103.30 120.00

101.10 Statistics r
100.00 120.00 (LCF to UTF) Volcan. Laye 230.00
a) max. number of micro- and macrocycles

undated
b) min. number of micro- and macrocycles

98.50
? a) maximum number of 60.00
r
MTP cycles Volcan. Laye
?

-4
- microcycles: 738
93.90
93.30 - macrocycles: 176 undated
110.00
?
b) minimum number of
cycles 220.00
?

90.20 110.00 Volcaniclastic Layer


- microcycles: 629
- macrocycles: 144 undated
50.00
86.70
c) average thickness
Volcaniclastic Layer
- microcycle (min.-
max.): 0.62-0.73 m 100.00 undated
- macrocycle (min.-
max.): 2.59-3.17 m
80.30 210.00
100.00 (including projections
from reference section
Latemartürme) 40.00

90.00
72.20
Lower Cyclic Facies ( LCF )

200.00
69.80 90.00

L-10 30.00
Lower Cyclic Facies ( LCF ) sensu Goldhammer

L-11

63.10 80.00

190.00
80.00
Volcanic lastic Layer
58.50
undated 20.00
Lower Cyclic Facies ( LCF )

70.00 volcaniclastic layer


52.60 crystal tuff
Lower Cyclic Facies ( LCF )

early middle Triassic / “Anisian”

70.00 base
180.00 undated
wind2
Lower Cyclic Facies ( LCF )

47.70 10.00
Secedensis Zone / late “Illyrian”

42.90 60.00
Lower Cyclic Facies (LCF )

170.00
60.00
39.00
0.00

Volcaniclastic Layer
33.80
50.00 volcaniclastic layer
undated
Statistics 160.00
50.00
(LCF to UTF)
a) max. number of micro- and macrocycles

b) min. number of micro- and macrocycles

a) maximum number of
cycles
40.00
- microcycles: 660
22.60 - macrocycles: 160
b) minimum number of 150.00
40.00 cycles
L-11 - microcycles: 579
- macrocycles: 136
L-1
c) average thickness
13.60 - microcycle (min.-
30.00
max.): 0.70-0.80 m
- macrocycle (min.-

LAT-31
max.): 2.90-3.14 m 140.00
30.00
(including projections
tic Layer
Volcaniclas from reference section
Latemartürme)

20.00

130.00
0.00 20.00
LAT-31
base
wind1
242.6 +0.7
- 0.7 Ma
Low. Tepee F. (LTF)

Statistics U-Pb isotopic age from single


10.00 zircons in volcaniclastic layer
(LCF to UCF sensu Mundil et al. (2003)
Goldhammer 1987,
Goldhammer et al. 120.00 Valsordaspitze, sampled
(1989, 1993) 10.00
Low. Tepee F. (LTF)

Level 28.60 m
b) alternative cyclostratigraphy (based on thinning upward trends)

Erzlahnspitze
Lower Tepee Facies ( LTF )

volcaniclastic layer below


a) cyclostratigraphy according to GOLDHAMMER (1987)

- number of 5th order


precessional cycles base of section
(microcycles) Latemartürme, projected
488 0.00 along traced physical surface
LTF

Level 29.00 m
- number of 4th order
eccentricity cycles 110.00 Valle della Veccia
0.00
(macrocycles) volcaniclastic layer below
97 base of section

- average thickness of Statistics Spiz de la Cagnota


precessional cycles (LCF to UTF) volcaniclastic layer not
preserved or deposited
a) max. number of micro- and macrocycles

(microcycles)
b) min. number of micro- and macrocycles

Statistics
0.66 m
a) max. number of micro- and macrocycles

(LCF to UTF) a) maximum number of


cycles
b) min. number of micro- and macrocycles

The uppermost 130 m of 100.00


the cyclic succession in - microcycles: 701
a) maximum number of - macrocycles: 167
the Latemar have not cycles
been measured or ana-
lyzed. - microcycles: 745 b) minimum number of
- macrocycles: 176 cycles
Goldhammer et al. - microcycles: 619
(1993) estimated, that b) minimum number of - macrocycles: 134
this part of the succes- cycles
sion comprises 110 cy- - microcycles: 621 c) average thickness
cles. - macrocycles: 148 90.00
- microcycle (min.-
They assumed, that the max.): 0.65-0.74 m
whole cyclic succession c) average thickness - macrocycle (min.-
is made up of “over 500 - microcycle (min.- max.): 2.73-3.40 m
m-scale fifth-order cy- max.): 0.64-0.76 m
cles” (1993:362), 598 cy- - macrocycle (min.- (including projections
cles (1993:363-365) max.): 2.69-3.20 m from reference section
Latemartürme)
(including projections
from reference section 80.00
Latemartürme)

? ?

70.00
Lower Platform Facies ( LPF )

early middle Triassic / “Anisian”


Lower Platform Facies ( LPF )

upper Reitzi Zone / late “Illyrian”

60.00

50.00

LPF 40.00

Lower
Platform
Facies 30.00

20.00

CHART 1.—Latemar cyclostratigraphy, biostratigraphy, and chronostratigraphy. Columns A3 to A7 show the


cyclostratigraphic sections measured in this study. Columns A1 and A2 illustrate the cyclostratigraphic 10.00

database by Goldhammer (1987; Appendix 4) and Preto et al. (2001; Data Repository, p. 2–3). Physically
traced and correlated macrocycle tops between the Valsordaspitze (A3) and Spiz de la Cagnota (A7)
sections show the internal architecture of the platform interior and the platform-wide persistence of
macrocycles. For technical reasons physically traced and correlated microcycle tops were not plotted. 0.00

Location Map Location Map True distances between vertical sections


4.31 km Legend / Latemar Sections
G OLDHAMMER (1987), PRETO et al. (2001) This study 1.57 km 0.75 km 0.94 km 1.05 km

B to Karer-Pass C to Karer-Pass D
Passo di Carezza Lithofacies type (LFT) 1: wacke- to grainstones marker horizons of Egenhoff et al. (1999)
Passo di Carezza UTF UTF R
Fault Fault Upper Teppee Facies (subtidal to shallow subtidal) (physically traced and correlated)
GOLD- N ? N
HAMMER slo cyclostratigraphic boundary
Section GOLD- Section pe Lithofacies type (LFT) 2: dolomitic caps
L-13 HAMMER Erzlahnspitze pla LPF-LTF-LCF-MTF-UCF-UTF
115 m Section 193 m tf. t (peri- to intertidal)
op
TF to UCF L-11 MTF to UCF
43 m macrocycle tops (physically traced and correlated)
to Ober-
eggen
LCF PRETO et al.
Section
to Ober- UCF UCF
160 m
eggen ? Upper Cyclic Facies Lithofacies type (LFT) 3: tepees
GOLD- MTF to UCF
? (inter- to supratidal) shallowing-/thinning-upward macrocycles and
Erzlahnspitze

Section
HAMMER
Latemartürme

GOLD- Valle della Vècia microcycles


Section HAMMER Section 232 m
L-FS II/I Biv. Sieff Biv. Sieff
87 m
Section Latemartürme
470 m
LCF to UCF Section Lithofacies type 4 (LFT): float-/rud-/bindstone
Valle de la Vècia

L-10 Spiz de la Cagnota


op

TF LTF to UTF associated with residual sediment base


e

92 m 316 m bases and tops of sliding windows (7) for spectral


s lo p

wind5
tf. t

LCF to TF LPF to MTF (inter- to supratidal) analyses in reference section Latemartürme


top
pla

Section Line of MTF Lithofacies type 5 (LFT): residual sediment wind3


Valsordaspitze
transect
?
Middle Tepee Facies MTF
Spiz de la Cagnota

C. dei Fuori to Forno (supratidal, reworked)


Valsordaspitze

sampling location for absolute age dating (vol-


p

229 m
to

LTF to MTF section name caniclastic layer) by Mundil et al. (2003)


pe
f.

thickness
at
slo

Rif. Torre ? stratigr. range Lithofacies type 6 (LFT): sheet cracks


pl

Section
di Pisa Cavignon (lateral tracing)
section name
Rif. Torre (associated with 2,5 and 3,4) Sections are leveled to the boundary MTF-UCF. Dolomitic
GOLD-
thickness
stratigr. range di Pisa
56 m
LTF to LCF
platform top to
slope boundary
LCF LCF
HAMMER Lower Cyclic Facies caps, supratidal limestones, residual sediments and sheet
Section
L-1
0 [km] 0.5 0 [km] 0.5 Lithofacies type 7 (LFT): volcaniclastics cracks in cycles thinner than 0.15 m are not represented (black
to Alpe di to Alpe di
Pampeago 23 m Pampeago ? (primary deposits from as clouds) line).
LCF
Contour Interval: 25 m Contour Interval: 25 m
LTF LTF
v04/06/01

Das könnte Ihnen auch gefallen