ANNUAL REVIEW MEETING October 15 - 16, 2012

Transcription

ANNUAL REVIEW MEETING October 15 - 16, 2012
UNIVERSITY OF MIAMI
ROSENSTIEL
SCHOOL of MARINE &
ATMOSPHERIC SCIENCE
CSL
CENTER FOR CARBONATE RESEARCH
ANNUAL REVIEW
MEETING
October 15 - 16, 2012
CSL
CENTER FOR CARBONATE RESEARCH
INDUSTRIAL ASSOCIATES
CSL – CENTER FOR CARBONATE RESEARCH
ANNUAL REVIEW MEETING 2012
TABLE OF CONTENTS
Meeting Program ................................................................................................... iii
Meeting Participants .............................................................................................. vii
Members & Associates ............................................................................................. x
Sea-Level Oscillations during the Sea-Level Highstands: Evidence from
the Basinal Section off Great Bahama Bank Gregor P. Eberli.......................................................................................................................................... 1
Evidence and Amplitude of Sea-Level Oscillations during the Last
Interglacial Highstand (MIS 5e) from the Bahamas Kelly L. Jackson, Gregor P. Eberli, Samuel B. Reid, Donald F. McNeill, and Paul M. Harris............. 5
Late Quaternary Growth History of Glover’s Reef, Belize: Insights into
Reef Distribution and Quaternary Sea Level Noelle J. Van Ee, Gregor P. Eberli, Flavio S. Anselmetti, Peter K. Swart, and
(EHUKDUGGischler .................................................................................................................................... 11
Influence of Depositional and Diagenetic Heterogeneities on Hydraulic
Conductivity in Reefal Carbonates: Preliminary Packer Injection Tests in
the Southern Dominican Republic Viviana Díaz, James S. Klaus, Donald F. McNeill, and Peter K. Swart ............................................... 17
Origin and Diagenesis of Microbialites on the Uplifted Atoll of Maré, New
Caledonia Chelsea L. Pederson, Donald F. McNeill, James S. Klaus, and Peter K. Swart .................................... 21
Microbial Community Characterization and Functional Gene Diversity of
Oolitic Grains from Great Bahama Bank Mara R. Diaz, Alan M. Piggot, Gregor P. Eberli, and James S. Klaus ................................................. 25
Pore Structure and Petrophysical Characterization of Microbialites Gregor P. Eberli, Ralf J. Weger, Jan Norbisrath, and Giovanna della Porta ................................... 29
Rock Fluid Interaction: How Dissolution Induced Changes in Pore
Structure Affect Acoustic Velocity Ralf J. Weger, Peter K. Swart, Gregor P. Eberli, and Mark Knackstedt ..............................................33
New Insights into Slope Processes from the Bahamas and West Florida Gregor P. Eberli, Donald F. McNeill, Thierry Mulder, Emanuelle Ducassou, Dierk Hebblen,
Claudia Wienberg, and Paul Wintersteller .......................................................................................... 37
Variability of Slope and Basin Floor Morphology along Southwestern
Great Bahama Bank Andrew Jo, Gregor P. Eberli, and Mark Grasmueck .............................................................................43
Composition of Cold-Water Coral Mound “Matterhorn” and its
Surrounding Sediments in The Straits of Florida Rani Sianipar, Gregor P. Eberli, and Emmanuelle Ducasou .............................................................. 49
i
Petrophysical Perspective of Cretaceous-Tertiary Re-deposited
Carbonates From the Apennines and the Adriatic Sea, Italy Irena A. Maura, Gregor P. Eberli, and Daniel Bernoulli ..................................................................... 57
Sedimentology, Geometries and Link to the Subsurface from a Field-Scale
Analog: The Sierra de la Vaca Muerta Michael Zeller, Samuel B. Reid, David L. Giunta, Ralf J. Weger, Gregor P. Eberli, and
-RVHLuis Massaferro ..............................................................................................................................63
Decoupled Inorganic and Organic Carbon Isotope Records: A Global
Signal Unrelated to Global Carbon Cycling? Amanda M. Oehlert and Peter K. Swart ................................................................................................. 73
Application of Cavity Ringdown Spectroscopy to Stable Isotopic
Monitoring of CO2 Sequestration during Enhanced Oil Recovery Benjamin T. Galfond, Daniel D. Riemer, and Peter K. Swart ............................................................... 79
Sub-Micron Digital Image Analysis (BIBSEM-DIA), Pore Geometries and
Electrical Resistivity in Carbonate rocks Jan H. Norbisrath, Gregor P. Eberli, Ralf J. Weger, Klaas Verwer, Janos Urai,
*XLOODXPHDesbois, and Ben Laurich .................................................................................................... 83
Using Clumped Isotopes to Understand Early Diagenesis Peter K. Swart, Monica M. Arienzo, Sean T. Murray, Yula Hernawati, James S. Klaus, and
Donald F. McNeill .................................................................................................................................... 91
New Insights into Dolomitization Using Clumped Isotopes Sean T. Murray, Monica M. Arienzo, and Peter K. Swart ....................................................................95
Speleothems: A Model System for the Study of Fluid Inclusions and
Clumped Isotopes Monica M. Arienzo, Sean T. Murray, Hubert B. Vonhof, and Peter K. Swart ................................... 101
Seismic and GPR Imaging of Fractures in Carbonate Reservoirs Using 3D
Diffraction Responses Caused by Fracture Intersections Mark Grasmueck, Tijmen Jan Moser, and Michael A. Pelissier....................................................... 107
4D GPR for Characterization of Fluid Flow in Carbonates: Insights from
Structural- vs. Stratigraphic-Controlled Domains and Comparison with
Eclipse Dynamic Modeling Pierpaolo Marchesini, Mark Grasmueck, Gregor P. Eberli, and Ralf J. Weger ................................ 113
ii
MEETING PROGRAM
2012 Annual Review Meeting
MONDAY
OCTOBER 15, 2012
08:30
Coffee
09:00
Welcome and Introduction
09:10
Sea-Level Oscillations during the Sea-Level Highstands:
Evidence from the Basinal Section off Great Bahama Bank
MORNING
Gregor P. Eberli
Gregor P. Eberli........................................................................................................................ 1
09:30
Evidence and Amplitude of Sea-Level Oscillations during
the Last Interglacial Highstand (MIS 5e) from the
Bahamas
Kelly L. Jackson, Gregor P. Eberli, Samuel B. Reid, Donald F. McNeill, and
Paul (Mitch) Harris.................................................................................................................. 5
09:50
Late Quaternary Growth History of Glover’s Reef, Belize:
Insights into Reef Distribution and Quaternary Sea Level
Noelle J. Van Ee, Gregor P. Eberli, Eberhard Gischler, and Flavio Anselmetti ................. 11
10:10
Coffee Break
10:40
Influence of Depositional and Diagenetic Heterogeneities
on Hydraulic Conductivity in Reefal Carbonates:
Preliminary Packer Injection Tests in the Southern
Dominican Republic
Viviana Diaz, James S. Klaus, Donald F. McNeill, and Peter K. Swart ............................. 17
10:55
Origin and Diagenesis of Microbialites on the Uplifted
Atoll of Maré, New Caledonia
Chelsea Pederson, Donald F. McNeill, James S. Klaus, and Peter K. Swart ...................... 21
11:15
Bacterial Community Characterization and Functional
Gene Diversity of Oolitic Grains from Great Bahama Bank
Mara R. Diaz, Alan M. Piggot, Gregor P. Eberli, and James S. Klaus ............................... 25
11:30
Pore Structure and Petrophysical Characterization of
Microbialites
Gregor P. Eberli, Ralf J. Weger, and Giovanna della Porta .............................................. 29
11:45
Rock Fluid Interaction: How Dissolution Induced Changes
in Pore Structure Affect Acoustic Velocity
Ralf J. Weger, Peter K. Swart, Gregor P. Eberli, and Mark Knackstedt ....................33
12:00
Lunch
iii
MONDAY
OCTOBER 15, 2012
13:15
New Insights into Slope Processes from the Bahamas and
West Florida
AFTERNOON
Gregor P. Eberli, Donald F. McNeill, Thierry Mulder, Emanuelle Ducassou,
Dierk Hebblen, Claudia Wienberg, and Paul Wintersteller ............................................... 37
13:45
Variability of Slope and Basin Floor Morphology along
Southwestern Great Bahama Bank
Andrew Jo, Gregor P. Eberli, and Mark Grasmueck ...........................................................43
14:05
Composition of the Cold-Water Coral Mound
“Matterhorn” and It’s Surrounding Sediments in the
Straits of Florida
Rani Sianipar, Gregor P. Eberli, and Emanuelle Ducassou............................................... 49
14:20
Petrophysical Perspective of Cretaceous-Tertiary Redeposited Carbonates from The Apenninnes and The
Adriatic Sea, Italy
Irena A. Maura, Gregor P. Eberli, and Daniel Bernoulli .................................................... 57
14:40
Coffee Break
15:15
Sedimentology, Geometries and Link to the Subsurface
from a Field-Scale Analog - The Sierra de la Vaca Muerta
Michael Zeller, Samuel B. Reid, David L. Giunta, Ralf J. Weger,
Gregor P. Eberli, and Jose Luis Massaferro ........................................................................63
15:45
Decoupled Inorganic and Organic Carbon Isotope
Records: A Global Signal Unrelated to Global Carbon
Cycling?
Amanda Oehlert and Peter K. Swart .................................................................................... 73
16:05
Application of Cavity Ringdown Spectroscopy to Stable
Isotopic Monitoring of CO2 Sequestration during
Enhanced Oil Recovery
16:20
Sub-Micron Digital Image Analysis (BIBSEM-DIA), Pore
Geometries and Electrical Resistivity in Carbonates
Benjamin T. Galfond, Daniel D. Riemer, and Peter K. Swart ............................................. 79
Jan H. Norbisrath, Gregor P. Eberli, Ralf J. Weger, Klaas Verwer, Janos Urai,
Guillaume Desbois, and Ben Laurich ................................................................................... 83
16:45
Reception in Commons
18:15
Return to Hotel
iv
TUESDAY
OCTOBER 16, 2012
08:00
Coffee
08:15
Using Clumped Isotopes to Understand Early Diagenesis
08:35
New Insight into Dolomitization using Clumped Isotopes
08:50
Speleothems: A Model System for the Study of Fluid
Inclusions and Clumped Isotopes
MORNING
Peter K. Swart, Monica M. Arienzo, Sean T. Murray, Yula Hernawati,
James S. Klaus, and Donald F. McNeill ................................................................................ 91
Sean T. Murray, Monica M. Arienzo, and Peter K. Swart ..................................................95
Monica M. Arienzo, Sean T. Murray, Hubert B. Vonhof, and Peter K. Swart ................ 101
09:10
Coffee Break
09:50
Seismic and GPR Imaging of Fractures in Carbonate
Reservoirs using 3D Diffraction Responses Caused by
Fracture Intersections
10:10
4D GPR for Characterization of Fluid Flow in Carbonates:
Insights from Structural- versus Stratigraphic-controlled
Domains and Comparison with Eclipse Dynamic Modeling
Mark Grasmueck, Tijmen Jan Moser, and Michael A. Pelissier ....................................... 107
Pierpaolo Marchesini, Mark Grasmueck, Gregor P. Eberli, and Ralf J. Weger .............. 113
10:30
Future Projects & Discussion
12:15
Lunch
v
TUESDAY
OCTOBER 16, 2012
13:15
Introduction to Fieldtrip
13:45
Core Workshop: Reef Successions in the Dominican
Republic and Pleistocene Oolitic Grainstones
15:00
Fieldtrip departs
vi
AFTERNOON
INDUSTRIAL ASSOCIATES ANNUAL REVIEW MEETING
CSL - CENTER FOR CARBONATE RESEARCH
MEETING PARTICIPANTS
Miami, October 15 - 16, 2012
AbdulJaleel AbuBshait
Saudi Aramco
EXPEC Advanced Research Center
Expec Annex Bldg. 137, Rm. 2660
Dhahran 31311, Saudi Arabia
[email protected]
Marcelo Blauth
PETROBRAS
E&P/ENGP/PR/GR
Av. Republica do Chile, 330
9 Andar – Centro
20031-170 Rio de Janeiro, RJ, Brazil
[email protected]
Frederico Andrade
Petrobras S.A.
UO-BC/ATP-NE/RES
Edifício geofísico Caseli, 1 andar
Avenida Elias Agostinho, 665
ZIP 27913- 350 - Imbetiba
Macaé, RJ, Brazil
[email protected]
Ornella Borromeo
ENI S.p.A.
Exploration & Production Division
Via Emilia 1
20097 S. Donato Milanese, Italy
[email protected]
Jiro Asada
Inpex Corporation
Akasada Biz Tower, 5-3-1-Akasaka
Minato-ku, Tokyo 107-6332
Japan
[email protected]
Vivek Chitale
BP America Inc.
501 Westlake Park Blvd.
Houston, TX 77079
[email protected]
Gregor Bächle
Shell International Exploration and
Production Inc.
200 North Dairy Ashford, WCK5328
Houston, TX 77079, USA
[email protected]
Anita Csoma
ConocoPhillips
600 N. Dairy Ashford, PR 3064
Houston, TX 77079
[email protected]
Alexandre Berner
Petrobras S.A.
UN-RIO/ATP-BRC/RES
Avenida Presidente Vargas 3131/1402
Rio de Janeiro, RJ, Brazil
Zip Code 20210-911
[email protected]
Alfredo Grell
Petrobras S.A.
E&P/UO-BC / ATP-C-S / RES
Av. Elias Agostinho, 665 - Imbetiba
Zip code:27913 350
Macaé, RJ, Brazil
[email protected]
vii
Paul Gucwa
Bahamas Petroleum Company
Montague Sterling Center, 2nd Floor
East Bay Street
P.O. Bxo SS-6276
Nassau, Bahamas
[email protected]
Jeroen Kenter
Statoil ASA
Sandsliveien 90
5254 Sandsli, Bergen, Norway
[email protected]
Mitch Harris
Chevron Energy Technology Company
6001 Bollinger Canyon Rd., D-1212
San Ramon, CA 94583-2324
[email protected]
Jesse Koch
BP America
501 Westlake Park Blvd.
Houston, TX 77079
[email protected]
Claude-Alain Hasler
Shell Global Solutions
Kessler Park 1
2288 GS Rijswijk
The Netherlands
[email protected]
Jose Luis Massaferro
YPF S.A.
Macacha Güemes 515 (C1106BKK),
Puerto Madero
Buenos Aires, Argentina
[email protected]
Karl Henck
BP – Brazil
North Campos Exploration TL
501 WestLake Park Blvd.
Houston, TX 77079
[email protected]
Paul Milroy
BG Group Exploration & Production
Thames Valley Park
Reading RG6 1PT
United Kingdom
[email protected]
Peter Holterhoff
Hess Corporation
Hess Tower
1501 McKinney
Houston, TX 77010
[email protected]
Elena Morettini
YPF S.A.
Piso 23
Macacha Güemes 515 (C1106BKK),
Buenos Aires, Argentina
[email protected]
Iulian Hulea
Shell Global Solutions
Postbus 60
2280 AB Rijswijk
The Netherlands
[email protected]
Leo Piccoli
BP Exploration & Production Inc.
580 Westlake Park Blvd.
Houston, TX 77079
[email protected]
viii
Chris Piela
BP America Inc.
501 WestLake Park Blvd.
Houston, TX 77079
[email protected]
Ingo Steinhoff
BP America Inc.
200 Westlake Park Blvd
Houston, TX 77079
[email protected]
Aimee Scheffer
ConocoPhillips
600 N. Dairy Ashford
Houston, TX 77079
[email protected]
Mario Suárez
Ecopetrol S.A.
Calle 37 No. 8-43 Edificio Colgás –
Piso 8
Bogotá, Colombia
[email protected]
Clara Sotelo
Ecopetrol S.A.
Calle 37 No. 8-43 Edificio Colgás –
Piso 8
Bogotá, Colombia
[email protected]
Alana Tischuk
Brazil Team
Maersk Oil Houston
2500 Citywest Blvd., Suite 2500
Houston, TX 77042-3041
[email protected]
Alice Stagner
ConocoPhillips
600 N. Dairy Ashford
Houston, TX 77079
[email protected]
Anette Uldall
Maersk Oil & Gas
Esplanaden 50
1263 Copenhagen K
Denmark
[email protected]
Carl Steffensen
BP America Inc.
200 WestLake Park Blvd.
Houston, TX 77079
[email protected]
Klaas Verwer
Statoil ASA
Sandsliveien 90
5254 Sandsli, Bergen, Norway
[email protected]
Mark Steinhauff
Saudi Aramco
EXPEC Advanced Research Center
Expec Annex Bldg. 137, Rm. 2670
Dhaharan 31311, Saudi Arabia
[email protected]
ix
CSL - CENTER FOR CARBONATE RESEARCH
MEMBERS & ASSOCIATES
Miami, October 15 - 16, 2012
PRINCIPAL INVESTIGATORS
Gregor P. Eberli
Mark Grasmueck
James S. Klaus
Donald F. McNeill
Peter K. Swart
Professor, Seismic Stratigraphy
Associate Professor, Subsurface Imaging
Assistant Professor, Paleontology
Scientist, Sedimentology
Professor, Geochemistry
ASSOCIATE SCIENTISTS
Mara R. Diaz
Greta Mackenzie
Ralf J. Weger
SCIENTIFIC COLLABORATORS
Emmanuelle Ducassou
Dierk Hebbeln
Mark A. Knackstedt
Thierry Mulder
Claudia Wienberg
University of Bordeaux, France
University of Bremen, Germany
Australian National University, Australia
University of Bordeaux, France
University of Bremen, Germany
VISITING RESEARCHER
Marcelo Blauth
Petrobras, S.A.
STUDENTS
Monica Arienzo, Deniz Atasoy, Caitlin Augustin, Quinn Devlin, Viviana Diaz, Ben
Galfond, Kelly Jackson, Andrew Jo, Deniz Kula, Pierpaolo Marchesini, Irena Maura,
Sean Murray, Jan Norbisrath, Amanda Oehlert, Erica Parke, Al Piggot, Rani Sianipar,
Hasan Calgar Usdun, Noelle Van Ee, Michael Zeller
RESEARCH ASSOCIATES
Amel Saied
STAFF
Karen Neher
Cory Schroeder
x
Office Manager
Technical Specialist
SEA-LEVEL OSCILLATIONS DURING THE SEA-LEVEL
HIGHSTANDS: EVIDENCE FROM THE BASINAL SECTION
OFF GREAT BAHAMA BANK
Gregor P. Eberli
KEY FINDINGS
ƒ
Cores retrieved during ODP Leg 166 in the Straits of Florida recovered a thick
succession of alternating meter-thick, marl-rich, pelagic/neritic limestone
alternations (“cycles”) that are largely paced by the orbital precession.
ƒ
Several intervals display, however, a sub-orbital frequency in the alternations:
ƒ
o
The interval between 12.7-13.6 myrs (802-910 mbsf) has
marl/limestone alternations, indicating a cycle duration of 11.1 kyrs.
81
o
Spectral analysis of the gamma log of a Late Miocene interval shows a
strong peak of 11 kyrs, in addition to the orbital frequencies.
o
In the early Pliocene, the į18O record of the shallow-dwelling foraminifera
G. sacculifer and the aragonite content are dominated by subMilankovitch variability.
Sub-orbital climate and sea-level changes occurred throughout the Neogene.
SIGNIFICANCE
Carbonate cyclostratigraphy generally assumes that the highest frequency of climate
driven sea-level changes are produced by insulation changes produced by the orbital
precession, i.e. 19-21 kyrs. In many modeling efforts of the carbonate system the timing
and frequency of sea-level fluctuations are, thus, computed within the Milankovitch
frequencies. In addition, spectral analyses of carbonate cycles regularly tune the time
series to one of the Milankovith frequency prior to the statistical tests.
There is, however, increasing evidence that oscillations during sea-level highstands
occur. If this is the case, a revisiting is necessary of the driving forces forming carbonate
cycles that are often reservoir flow units. To prove the sub-orbital frequency of sea-level
oscillations a precise age model is needed but precise dating is an inherent problem in
shallow-water carbonates. The periplatform area of Great Bahama Bank offers the
opportunity to assess the frequency of platform-derived material in well-dated sections.
SEDIMENTARY CYCLES IN THE STRAITS OF FLORIDA
The marl-limestone alternations (“ cycles”) in the Straits of Florida
On the slopes and in the basins surrounding the Great Bahama Bank, aragonite cycles
and turbidite composition are indicators of high-frequency sea-level fluctuations (e.g.,
Schlager et al., 1994). Cores retrieved during ODP Leg 166 in the Straits of Florida
recovered a thick succession of alternations (“cycles”) between meter thick layers with
platform-derived material and thin layers with more pelagic sediments (Figure 1).
1
Carbonate-rich intervals are interpreted to reflect periods of high sea level when the
platform is flooded and aragonite mud is exported to the basins while the thin intervals
correspond to times of increased pelagic and siliciclastic input during sea-level lowstands
and early transgression before the platform is flooded again (Eberli et al., 1997; Rendle
and Reijmer, 2002; Betzler et al., 1999; Isern and Anselmetti, 2002). The cycles can be
recognized in cores and on logs (Figure 2).
Figure 1. (Left) Schematic drawing of the marl-limestone alternations in the Straits of Florida. The white
limestone is composed of mostly platform derived-material with over 90% aragonite in the Pleistocene
sections. The dark intervals are always thinner and are composed of pelagic foraminifers and nannofossils
and various amounts of silt and clay (modified from Eberli et al., 1997). (Right) Core photograph across
Miocene marl-limestone alternation with photomicrographs and SEM images documenting the
composition and diagenesis. Neritic components are visible in the light limestone portion while pelagic
foraminifers are seen on the photomicrographs and clay minerals in the SEM images of the dark intervals.
Because the limestone intervals contain platform-derived material, they are interpreted to represent times
of high sea level.
Based on coupled bio- and cyclostratigraphy Kroon et al. (2000) determined that these
cycles are tied to orbital forcing mechanisms. Using the Formation MicroScanner (FMS)
logs to measure cycle thickness Williams et al. (2002) show that the sedimentary cycles
are paced by the Earth’s climatic precession for the time interval of 9-12 myrs. In
addition, the cycle thickness contains long-term cycles that repeat about every 400 kyrs,
which they correlate to the 400 kyrs cycles in orbital eccentricity.
Evidence of sub-orbital cycles
Two studies found sub-orbital signals in the marl-limestone alternations. Bernet (2001)
documents cycles of 11 kyrs duration in two intervals of along the Bahamas Transect. The
first is between 802-910 m at ODP site 1006 and is dated 12.7-13.6 myrs. Within this 900
kyrs time interval are 81 marl-rich pelagic/neritic limestone alternations, indicating
2
cycle duration of 11.1 kyrs. In an older section, spectral analysis of the gamma-ray logs
within Upper Miocene marl/limestone alternations in Late Miocene strata at ODP site
1003 produced strong peaks in orbital frequencies of obliquity (40 kyrs) and precession
(23 kyrs) but also a peak at 11kyrs that is stronger than the obliquity peak (Figure 3).
Figure 2. (Left) Core photograph of Miocene marl-limestone alternation in Site 1007, Core 1007 84R
(modified from Bernet, 2001). (Right) Cyclic variations in the FMS (Formation MicroScanner) resistivity
image log, FMS resistivity average, resistivity log (SFLU), and normal and high resolution porosity logs
(APLC,HALC) of the Pliocene section in Site ODP 1006. Picked cycles, used for cycle counting, are indicated
by horizontal lines (modified from Williams et al., 2002).
Reuning et al. (2006) analyzed magnetic susceptibility, aragonite content and į18O
records from two different planktonic foraminifera species in an early Pliocene core
interval from ODP site 1006. They found that the į18O record of the shallow-dwelling
foraminifera G. sacculifer and the aragonite content are dominated by sub-Milankovitch
variability. Because the magnetic susceptibility and the deeper-dwelling foraminifera G.
menardii show precession frequency they interpret the sub-Milankovitch frequency as a
climate rather than sea-level driven signal. The sub-orbital signal in the aragonite
content is, however, indicating that sea-level changed with the climate signal. Many
studies document that aragonite is produced on the platform during high sea level and
exported to the basins during these times (Boardman and Neumann, 1984; Schlager et
al., 1994; Rendle and Reijmer, 2002).
3
IMPLICATIONS
The sub-orbital signal in the marl/limestone alternations in the cores of ODP Leg 166
indicates that sub-orbital sea-level oscillations occurred throughout the Neogene. If
indeed sub-orbital oscillations are able to produce shallowing-upward carbonate cycles,
the possibility of tracing the orbital forcing mechanisms past the Mes0zoic is diminished
because large uncertainties are introduced in spectral analyses of these ancient deposits.
Sub-orbital sea-level oscillations might explain the overabundance of cycles in some
ancient successions, like the Latemar where roughly 600 cycles were deposited in 3
myrs.
Figure 3. Spectral analysis of
the Late Miocene strata using
the gamma ray log. In
addition to the obliquity and
the precession a peak at 11
kyrs indicates sub-orbital
cyclicity
(modified
from
Bernet, 2001).
REFERENCES
Bernet, K. H., 2001, The record of hierarchies of sea level fluctuations in cores, logs, and seismic
data along the Great Bahama Bank Transect: University of Miami Ph.D. Dissertation, 210 pp.
Betzler, C., Reijmer J. J. G., Bernet, K., Eberli, G. P., and F. S. Anselmetti, 1999, Sedimentary
patterns and geometries of the Bahamian outer carbonate ramp (Miocene and lower Pliocene,
Great Bahama Bank): Sedimentology, v. 46, p. 1127-1145.
Boardman, M. R., and A. C. Neumann, 1984, Sources of periplatform carbonate: Northwest
Providence Channel, Bahamas: Journal of Sedimentary Petrology, v. 50, p. 1121-1148.
Eberli, G. P., Swart, P. K., Malone, M., et al., 1997, Proceedings ODP, Init. Repts., 166: College
Station, TX (Ocean Drilling Program), 850 pp.
Kroon, D., Williams, T., Pirmez, C., Spezzaferri, T., Sato, T., and J. D. Wright, 2000, Coupled
early Pliocene- middle Miocene bio-cyclostratigraphy of Site 1006 reveals orbitally induced
cyclicity patterns of Great Bahama Bank carbonate production: Proceeding of ODP, Scientific
Results, v. 166, p. 155-166.
Rendle, R. H., and J. J. G. Reijmer, 2002, Quaternary slope development of the western, leeward
margin of Great Bahama Bank: Marine Geology, v. 185, p. 143-164.
Reuning, L., Reijmer, J. J. G., Betzler, C., Timmermann, A., and S. Steph, 2006, SubMilankovitch cycles in periplatform carbonates from the early Pliocene Great Bahama Bank:
Paleoceanography, v. 21, p. 1007-1021.
Schlager, W., Reijmer, J. J. G., and A. W. Droxler, 1994, Highstand shedding of carbonate
platforms: Journal of Sedimentary Research, v. B64, p. 270-281.
Williams, T., Kroon, D., and S. Spezzaferri, 2002, Middle-Upper Miocene cyclostratigraphy of
downhole logs and short to long term astronomical cycles in carbonate production of Great
Bahama Bank: Marine Geology, v. 185, p. 75-93.
4
EVIDENCE AND AMPLITUDE OF SEA-LEVEL OSCILLATIONS
DURING THE LAST INTERGLACIAL HIGHSTAND (MIS 5E)
FROM THE BAHAMAS
Kelly L. Jackson, Gregor P. Eberli, Samuel B. Reid, Donald F. McNeill,
and Paul M. Harris1
Chevron Energy Technology Company
KEY FINDINGS
ƒ
Positions of reefs, beach, and eolian deposits provide evidence that sea level
during the last interglacial highstand (MIS 5e) 115-125 ka was not a single rise
and fall but fluctuated with a minimum oscillation of 10 m within a few thousand
years.
ƒ
During MIS 5e, sea-level first rose to 7 m above present, then dropped by ~ 2 m
before the mid- MIS 5e sea-level drop of at least 10 m. Next, sea level rose to form
the younger MIS 5e highstand and then dropped at the end of MIS 5e in
downstepping pulses.
ƒ
Highstand sea-level oscillations create complex patterns of carbonate sediment
deposition and promote extensive lateral accretion.
IMPORTANCE
New evidence from New Providence Platform, Bahamas, indicates that sea-level
oscillated a minimum of 10 m during MIS 5e, creating early and late substages within the
5e highstand. A 10 m + oscillation exposed Great Bahama Bank, creating two separate
depositional cycles within 10,000 years. This highstand oscillation requires a suborbital
forcing mechanism of much shorter duration than Milankovitch frequencies. This is
important because it contradicts the preconceived notion that precession is the
controlling factor of high-frequency sequences and the building blocks of carbonate
cycles which we assume are reservoir flow units. In addition, this new evidence
documents that rapid climate changes, which previously were thought to only take place
during glacial times, can take place during warm interglacial periods.
Figure 1. Three criteria are used to identify sea level in core and outcrop in New Providence and the Exuma
Cays, Bahamas: (1) The elevation of the transition from beach to eolian dunes, (2) The elevation of reefs
plus an assumed water depth of 2-5 m, and (3) Exposure horizons (calcretes, caliche crusts) separating
subtidal facies in core.
5
SEDIMENTOLOGIC INDICATORS FOR SEA-LEVEL AMPLITUDE
Sea level was 6-7 m higher than present during the last interglacial highstand 115-125
ka (MIS 5e). Exposure horizons and lithologic changes in cores and outcrops combined
with age dating in the Exuma Cays and New Providence, Bahamas, provide
sedimentologic and stratigraphic evidence of sea-level oscillations during MIS 5e.
The position of sea level during MIS 5e is assessed using three criteria: 1) The elevation
of the transition from beach to eolian dunes, in particular the first occurrence of
keystone vugs in the beach sediments, 2) The elevation of reefs plus an assumed water
depth of 2-5 m, and 3) Exposure horizons (calcretes, caliche crusts) separating subtidal
facies in core document the sea-level drop below the former sediment surface (Figure 1).
1. Beaches and Eolian Dunes
Carbonate eolian dunes form adjacent to the modern beach and start cementing quickly
after deposition. Unlike siliciclastic dunes, carbonate eolian dunes do not migrate
therefore one ridge of eolian dunes implies one sea-level position in a carbonate coastal
system. On New Providence, three low-relief beach ridges prograde toward the modern
shoreline. Using the lowest occurrence of keystone vugs as a sea-level proxy, the beach
ridges downstep from +7.6 to +7.0 to +5.1 m above present sea level over a lateral area of
~4 km (Garrett and Gould, 1984; Reid, 2010). In between the +7.0 m and +5.1 m beach
ridges, there is a calcrete separating subtidal facies in core indicating a drop in sea level
during MIS 5e. At two locations within the Exuma Cays Land and Sea Park, there are five
parallel north-south trending eolian dune ridges that were deposited during MIS 5e.
Beach facies in core are slightly inclined laminated grainstones with fenestral pores.
Eolian facies in core feature high-angle cross-bedding, meniscus calcite cement and/or
equant-bladed and dogtooth spar. Both facies have interparticle porosity and
intraparticle-moldic porosity where the peloids and ooids started to dissolve away
(Figure 2). The subsurface position of the beach to dune transition in cores WW1-3 and
BI1-OB1 deepens from west to east, indicating downstepping toward Exuma Sound as a
result of pulsed ice build-up towards the glacial period.
2. Reefs
The presence of depth-specific coral species (i.e., Acropora palmata, Acropora
cervicornis, etc.) in core and outcrops are good indicators of past sea level. Pleistocene
shallow-water reefs are present at +1.5 m above present sea level on Rocky Dundas
(dominated by Acropora cervicornis) in the north-central Exumas as well as on Darby
Island (mixed shallow-water species) in the southern Exumas indicating that sea level
was higher than today when the reef was alive. At Bitter Guana Cay, the shallow-water
reef facies are at approximately -1.8 to -2.8 m below present sea level. Halley et al. (1991)
dated a +1.5 m reef terrace on the leeward side of Norman’s Pond Cay as 117 ± 5 to 119 ±
4 ka. Open system U-Th dating of Montastraea faviolata at -3.5 m depth dated the reef
on Little Darby to 122 ± 1.32 ka old. Both dates confirm that these reefs were alive during
MIS 5e. Thompson et al. (2011) documented stacked MIS 5e reefs in San Salvador,
Bahamas, with an older MIS 5e reef at 123.5 ka and a younger MIS 5e reef at 119.5 ka,
separated by an unconformity that they interpreted as an exposure related to the midMIS 5e sea-level drop. In the Little Darby core a younger reef is recovered above the
dated reef. Unfortunately, the contact is not recovered in the core and the younger reef is
diagentically too altered to allow U/Th dating.
6
3. Exposure Horizons within Subtidal Deposits
The presence of an exposure horizon separating subtidal facies in core indicates a drop
in sea level during MIS 5e. On New Providence, subtidal burrowed and bedded
grainstones in short cores feature a cm-scale calcrete 20-30 cm within MIS 5e subtidal
deposits, documenting an intermittent lowering of sea level during MIS 5e. Exposure
horizons are also present in cores from the Exumas separating eolian and subtidal facies.
Figure 2. (A) Photomicrograph of peloidal oolitic grainstone featuring fenestral porosity typically found in
beach to eolian dune environments. (B) Peloidal oolitic grainstone with intraparticle to moldic porosity
and solution enhancement of the original fenestral pores. (C) Eolian peloidal oolitic grainstone featuring
dogtooth and meniscus calcite cement, interparticle porosity, and intraparticle porosity. (D) Close-up of a
partially dissolved ooid with equant dogtooth spar surrounding the grains.
Summary of the Sea-level Oscillations during MIS 5e
New evidence from New Providence and the Exuma Cays reveals that sea-level
oscillated a minimum of 10 m and created early and late substages of MIS 5e. New
Providence beach deposits at +7.6 m above present sea level represent the older peak of
MIS 5e sea level. A down-stepping beach ridge indicates a subsequent sea-level position
at +7.0 m. A calcrete in the subtidal deposits adjacent to the beach documents the midMIS 5e sea-level drop. In the Exumas, a calcrete associated with this fall separates
subtidal facies at -2.0 m. Sea-level rises again to form the younger MIS 5e highstand; this
rise is represented by a beach ridge at +5.1 m on New Providence Island and Exumas
reefs up to +1.5 m above modern sea level. Parallel down-stepping beach to eolian dune
transitions provide evidence for a pulsed down-stepping of sea level at the end of MIS 5e.
The lowest occurrence of this transition is approximately -12 m below present sea level.
7
Taking the lowest occurrence of calcretes that mark the mid-MIS 5e sea-level fall and
the highest beach elevation into account, the MIS 5e sea-level oscillated a minimum of 10
m (9 m plus at least 1 m to expose the strata) (Figure 3). This sedimentologic evidence
corroborates the 15 m estimate from Caribbean corals (Thompson and Goldstein, 2005;
Thompson et al., 2011). In addition, the downstepping pattern within and at the end of
MIS 5e documents pulsed changes of sea level during MIS 5e that likely coincides with
pulsed ice buildup.
Figure 3. (A) MIS 5e sea level from Thompson and Goldstein (2005). (B) MIS 5e sea level interpreted from
the facies successions in the Exumas and New Providence, Halley et al. (1991), Aurell et al. (1995), and Reid
(2010).
COMPLEX CARBONATE SEDIMENT DEPOSITION AND ACCRETION
The Exuma Cays and surrounding carbonate sand bodies of Great Bahama Bank
feature a stacked succession of carbonate grainstones that display complex patterns of
facies juxtapositions and lateral accretion. This heterogeneity is the direct product of sealevel oscillations during Holocene and Pleistocene sea-level highstands. The recognition
of these suborbital sea-level oscillations explains complicated facies patterns in shallowwater carbonate strata.
The numerous and varied islands of the Exuma Cays span 170 km NE-SW and 5-10 km
W-E and are composed of Holocene (<6,000 ybp), MIS 5e (~125,000 ybp), and older
Pleistocene strata. These stratigraphic units are laterally accreted forming the complex
island stratigraphy of the Exumas. Field mapping and ground-truthing of satellite
imagery of one key island (Hawksbill Cay) documents patterns of Holocene and
Pleistocene facies accretion laterally; 38% of the island is Pleistocene at the surface and
Holocene ridges were deposited around the Pleistocene topography in a complex fashion
(Harris and Ellis, 2009). Holocene ridges have elevations from near sea level to +12 m
8
while Pleistocene landforms have elevations from near sea level to +19 m. This complex
pattern of carbonate sediment deposition and accretion extends along the entire Exumas
windward margin. Antecedent Pleistocene topography directly affects the distribution of
shoals and underwater environments along the margin.
The stratigraphic architecture resulting from sea-level oscillations produces
heterogeneous grainstones with a wide range of grain sizes, sedimentary structures, and
diagenetic features. Many cores, including HY1 (Figure 4) feature selective cementation
or diagenetic partitioning of the oolitic peloidal grainstone. Common to most cores is the
presence of moldic porosity (most commonly oomoldic porosity). This diagenesis
creating the porosity complicates dating stratigraphic units as most of the original
material has been diagenetically altered or dissolved away.
Figure 4. (A, B) Photomicrographs showing an example of differential cementation in eolianite ooid
grainstones from Harvey Cays. (B) In some areas the grains are almost completely dissolved away
showing intraparticle-moldic porosity while in other areas, the grains are micritized but remain intact.
IMPLICATIONS FOR CYCLOSTRATIGRAPHY AND RESERVOIR HETEROGENEITY
The recognition of suborbital sea-level oscillations has two major implications. First,
sea-level falls within interglacials indicate that a mechanism exists that can produce sealevel fluctuations during times when the Earth is considered ice free, i.e., the greenhouse
world. This might explain the cyclic nature of Cretaceous and Triassic strata. The short
duration of these oscillations produces uncertainty regarding the commonly accepted
notion of Milankovitch cyclostratigraphy in carbonates.
Second, the combined product of high-frequency orbital sea-level changes and
suborbital oscillations is a complex lateral and vertical stratigraphic architecture that
juxtaposes grainstones deposited in different environments. Antecedent topography
from older sea-level oscillations and the repetition of sea-level oscillations during each
subsequent highstand yields a complex stratigraphic architecture of lateral accretion and
overstepping wedges. Suborbital sea-level oscillations have dramatic effects on carbonate
depositional environments and occur on timescales of just a few thousand years. As
illustrated by the prograding wedges on New Providence, the highstand oscillations
enable the lateral accretion of shoal systems that one single sea-level rise would not
achieve.
9
REFERENCES
Aurell, M., McNeill, D. F., Guyomard, T., and P. Kindler, 1995, Pleistocene shallowing-upward
sequences in New Providence, Bahamas: signature of high-frequency sea-level fluctuations in
shallow carbonate platforms: Journal of Sedimentary Research, v. B65, no. 1, p. 170-182.
Garrett, P., and S. J. Gould, 1984, Geology of New Providence Island, Bahamas: Geological
Society of America Bulletin, v. 95, p. 209-220.
Halley, R. B., Muhs, D. R., Shinn, E. A., Dill, R. F., and J. L. Kindinger, 1991, A +1.5-m reef terrace
in the southern Exuma Islands, Bahamas: Geological Society of America Abstracts and
Programs, v. 23, no. 1, p. 40.
Harris, P. M., and J. Ellis, 2009, Satellite imagery, visualization and geological interpretation of
the Exumas, Great Bahama Bank: an analog for carbonate sand reservoirs: SEPM Short
Course Notes No. 53, DVD.
Reid, S. B., 2010, The complex architecture of New Providence Island (Bahamas) built by multiple
Pleistocene sea level highstands: University of Miami M.S. Thesis, Open Access Theses, Paper
77, http://scholarlyrepository.miami.edu/oa_theses/77.
Thompson, W. G., and S. L. Goldstein, 2005, Open-system coral ages reveal persistent suborbital
sea-level cycles: Science, v. 308, p. 401-404.
Thompson, W. G., Curran, H. A., Wilson, M. A, and B. White, 2011, Sea-level oscillations during
the last interglacial highstand recorded by Bahamas corals: Nature Geoscience, v. 4., p. 684687.
10
LATE QUATERNARY GROWTH HISTORY OF GLOVER’S
REEF, BELIZE: INSIGHTS INTO REEF DISTRIBUTION AND
QUATERNARY SEA LEVEL
Noelle J. Van Ee, Gregor P. Eberli, Flavio S. Anselmetti1, Peter K. Swart,
and Eberhard Gischler2
1)
EAWAG, Sedimentology Group, Dübendorf, Switzerland
2) Goethe University, Frankfurt am Main, Germany
KEY FINDINGS
ƒ
ƒ
ƒ
Pleistocene reefs strongly influence modern reef distribution.
-
93% of seismically imaged modern patch reefs sit on antecedent
topographic highs.
-
Modern patch reefs are comprised mainly of in situ reef facies.
Unequal filling of accommodation space and asymmetric map facies typify
Quaternary platform development.
-
No indication in seismic data of tilting or faulting of Late Pleistocene
succession.
-
Modern and Pleistocene bathymetry relate to differential ability of facies
and hydrodynamic regime to fill accommodation space.
-
Satellite facies mapping and grain size analysis shows windward –
leeward asymmetry on both a platform and patch reef scale.
Quaternary sea-level oscillations have left strong isotopic and petrographic
signatures while forcing innovative attempts to date Late Pleistocene strata.
Figure 1. Study area with location of cores and seismic lines. Inset image modified from Google.
11
SIGNIFICANCE
The location of Glover’s Reef proximate the North America – Caribbean plate boundary
and along a known precipitation gradient make it an ideal candidate for studying
possible controls on reef distribution and platform development. Despite separation
from the mainland during the Quaternary, Glover’s Reef is a heterogeneous carbonate
system. Understanding Glover’s Reef can provide insights for predicting occurrence,
extent, and quality of carbonate reservoirs in reef mound – patch reef complex or
isolated platform depositional systems.
RESULTS
Reef Distribution
A single-channel seismic survey of over 100 km of grid lines in conjunction with N-S
and E-W rotary core transects (Figure 1) show no evidence of faults/folds controlling
patch reef locations. Rather, 93% of surveyed Holocene patch reefs are located on
Pleistocene topographic highs (Figure 2). In situ reef facies in patch reef cores indicate
that these highs are growth-induced: reefs sit on reefs (Figure 3).
Figure 2. Seismic data showing influence of antecedent topography and patch reef core (see Figure 1 for
location). Holocene reefs and modern sea bottom are traced in green. The Pleistocene top is traced with
red. Not every Pleistocene topographic high results in a Holocene patch reef, but the vast majority (93%) of
surveyed Holocene patch reefs are located on antecedent highs.
Platform Development
Although modern bathymetry deepens in the south of the platform, there is no evidence
to support a tilted Pleistocene surface. Maps constructed of the Pleistocene topography
from the seismic grid show a relatively flat surface except for locally beneath patch reefs
(Figure 4). These grid maps also indicate that observed modern lagoon bathymetry is
caused by a wedge of Holocene sediment that is thickest in the northeast corner of the
lagoon and thins southward. The northeast corner is the most windward corner of the
platform and the location of one of only several breaks in the reef-rimmed platform.
Inconsistencies in the location of Pleistocene horizons in the cores correlate to
antecedent topography and variable growth potential of facies. Patch reef cores sit on
antecedent highs while windward cores contain facies with fast-growing Acropora
species and early marine cementation, in contrast to leeward core G2 (Figure 3).
12
Satellite-based facies mapping, and modern grain size analyses also provide evidence
for platform development in response to a dominant northeast to southwest trade wind
regime. Asymmetric facies belts are apparent on both a platform and patch reef scale
(Figure 5). Grain size analyses of windward and leeward transects illustrate that median
grain size decreases steadily with distance from the windward margin in contrast to a
sharp jump at 800 m from the leeward margin. The poorest sorting is found closest to
the margin on the windward side but furthest from the margin on the leeward side.
Figure 3. Rotary cores collapsed along leeward – windward transect shown with lithologies. OR =
Oreaster Reef, AR = Aurelia Reef, SW = Southwest Caye, MC = Middle Caye, ER = East Rim, NR =
North Rim. Note that cores GR3, OR, and AR are from lagoon patch reefs.
Sea-level History
Negative carbon and positive oxygen stable isotope excursions suggest three exposure
events in the Pleistocene, including the Last Glacial Maximum. Multiple exposures to
meteoric water has leached almost all original aragonite from the sediment and left a
petrographic signature of vuggy dissolution and recrystallization to blocky calcite (Figure
6). Lack of appropriate U-Th dating material has led to exploration of new techniques of
amino acid racemization applied to coral geochronology. The two youngest Pleistocene
sequences (P1 & P2) are relatively close in age, perhaps even substages of Marine Isotope
Stage 5e. Holocene U-Th and C-14 ages of coral and peat fall along existing sea-level
curves for the western Caribbean and suggest an accumulation of organic material
around 8000-7500 years before present (BP) and initial coral growth shortly thereafter
at approximately 7000 years BP (Figure 7).
13
Figure 4. Horizons traced on seismic lines were interpolated to create maps of modern bathymetry, depth
to the top of the Pleistocene sequence, and the thickness of Holocene sediment. The thickest areas of
sediment correspond to breaks in the platform rim.
Figure 5. Asymmetric distribution of modern facies and grain size distribution. (A) Asymmetric facies are
apparent on both platform and patch reef scale. (B) Windward transect GR7 is very coarse and poorly
sorted near the platform margin. (C) Leeward transect GR11 is most poorly sorted away from the
platform margin.
14
Figure 6. Quaternary sea-level history. (A) Negative carbon and positive oxygen stable isotope excursions
are interpreted to represent exposure events in the Southwest Caye core. (B) Three values from two
Pleistocene sequences cluster together on a DL Glutamic – Aspartic acid cross plot, suggesting the
samples are close in age, perhaps even substages of MIS 5e. (C) Thin sections from the Southwest Caye
core display common petrographic signatures of meteoric diagenesis: blocky recrystallization to calcite
and vuggy dissolution.
INTERPRETATION AND IMPLICATIONS
Although previous work on the Belize margin has stressed that reef distribution is
controlled by underlying fault/fold structures or karst processes, we find that on Glover’s
Reef, Holocene reefs build on top of Pleistocene reefs. Many of the platform
heterogeneities can be explained by the strong control of consistent wind direction on
platform development. Additional vertical heterogeneities have been introduced by
platform response to high frequency sea-level oscillations. This work implies that even
isolated carbonate systems can build complicated architecture in relatively short time
intervals.
15
Figure 6. New U-Th and C-14 dates from Glover’s Reef compared to existing data from the
Western Atlantic.
REFERENCES
Blanchon, P., Jones, B., and D. C. Ford, 2002, Discovery of a submerged relic reef and shoreline
off Grand Cayman: further support for an early Holocene jump in sea level: Sedimentary
Geology, v. 147, p. 253-270.
Gischler, E., and J. H. Hudson, 1998, Holocene development of three isolated carbonate
platforms, Belize, Central America: Marine Geology, v. 144, p. 333-347.
Lightly, R.G., MacIntyre, I.G., and R. Stuckenrath, 1982, Acropora palmata reef framework: A
reliable indicator of sea level in the Western Atlantic for the past 10,000 years: Coral Reefs, v.
1, p. 125-130.
Toscano, M.A., and I. G. MacIntyre, 2003, Corrected Western Atlantic sea-level curve for the last
11,000 years based on calibrated C14 dates from Acropora palmata framework and intertidal
mangrove peat: Coral Reefs, v. 22, p. 257-270.
16
INFLUENCE OF DEPOSITIONAL AND DIAGENETIC
HETEROGENEITIES ON HYDRAULIC CONDUCTIVITY IN
REEFAL CARBONATES: PRELIMINARY PACKER INJECTION
TESTS IN THE SOUTHERN DOMINICAN REPUBLIC
Viviana Díaz, James S. Klaus, Donald F. McNeill, and Peter K. Swart
KEY FINDINGS
ƒ
Stratal packer well injection tests reveal a general trend of decreasing hydraulic
conductivity (K) measured over a 25 m interval that spans two Pleistocene highstand reef sequences.
ƒ
Decreasing (K) is related to depositional facies as well as diagenetic overprint
associated with dissolution and cementation in the meteoric environment.
ƒ
Hydraulic conductivity calculated from injection tests is, with exceptions,
positively correlated to plug permeability measurements and inversely correlated
to core recovery.
SIGNIFICANCE
Depositional and diagenetic heterogeneities within carbonate rocks can influence flow
and transport parameters, and it is increasingly recognized that sedimentological and
stratigraphic models can provide a method of incorporating geological variability into
models of fluid flow in both sedimentary aquifers (Fraser and Davis, 1998) and
petroleum reservoirs (Fogg and Lucia, 1990; Lucia, 2007). In carbonates, the porosity
and permeability structure is dependent on both matrix porosity and the development of
larger scale secondary porosity. The resulting complex porosity distribution in
carbonates is reflected in the hydraulic-conductivity. If one views dissolution zones as
discrete heterogeneities, the challenge of predicting transport in carbonate rocks is one
of characterizing the hydraulic-conductivity distribution at a scale that captures the
variability of these heterogeneities.
An ongoing challenge in assessing reservoir properties is the integration of multiple
data sources (thin section, core data, well log, geological models, seismic data, well tests,
simulation cells) and their variable scales (Figure 1). This study will integrate surface
and subsurface stratigraphic, depositional, diagenetic and petrophysical data with
hydrogeological data in order to characterize the factors that influence hydraulic
properties of Plio-Pleistocene reefal limestones of the Dominican Republic (Figures 2-3).
When completed, vertical profiles of hydraulic conductivity will be obtained from shortinterval packer tests performed in a transect of six wells drilled perpendicular to the
prograding depositional packages. Previous stratigraphic analyses of depositional and
diagenetic facies, combined with detailed petrophysical characterizations will be
combined with the short-interval (<1 m) packer tests. These packer tests provide for
calculation of hydraulic conductivity data on both matrix and dissolution zones. The
long-term goal of the project is to obtain hydrostratigraphic data from the PlioPleistocene carbonates that can ultimately be incorporated into a reservoir flow
simulation model.
17
Figure 1. A range of scales for data used in geological models and flow simulation models. The need for
upscaling and downscaling of porosity and permeability data versus coarser-scale facies and seismic data.
Figure 2. Location of Southern Dominican Republic. Cores 1 through 5 were drilled in a transect from the
oldest terrace at 50 m to the youngest terrace at 5 m above sea level.
18
Figure 3. Cross-section of prograding Plio-Pleistocene reefs in Southern Dominican Republic.
Figure shows core location and shallowing upward cycles with facies distribution.
RESULTS FROM THE FIELD TEST
Preliminary short interval (1 m) packer tests were performed in core 3 (30 m terrace)
from the surface to a depth of ~24 m. This interval is characterized by three distinct
highstand reef sequences separated by a subaerial exposure at ~12.5 m in the core. The
lower sequence is characterized by reef front facies dominated by branching corals in a
packstone matrix, the next sequence is characterized by a shallower reef crest facies with
a mix of massive and branching corals in a packstone to grainstone matrix. The
uppermost sequence is characterized by the back-reef lagoon with branching corals in a
wackestone to packstone matrix. Carbon and oxygen isotopes through this interval
reflect progressive diagenetic overprinting in the lower sequence. The light values of
Figure 4. Plot showing hydraulic conductivity with depth in core 3. Measurements span two highstand
reef sequences separated by a subaerial exposure surface. Stable isotope data and permeability data
from petrophysical cores is included for comparison.
19
carbon at the top of the core and at 12.5 m reflect the two periods of subaerial exposure.
Permeability values from petrophysical plugs ranged from 0.2-414.8 mD in the lower
interval, and 119.7-2850 mD in the upper interval. Values of hydraulic conductivity from
the 1m packer tests ranged between 0.007-0.013 cm/s in the lower unit and 0.011-0.024
cm/s in the upper unit. A general trend of decreasing hydraulic conductivity can be seen
from the surface to the subaerial exposure at 12.5 m (Figure 3). Core recovery, used as a
porosity proxy due to the absence of log data, appears inversely correlated to the
hydraulic conductivity. While plug permeability and hydraulic conductivity are
positively correlated below 5 m, they appeared inversely correlated in the upper 5 m of
the core.
INTERPRETATION AND IMPLICATIONS
Changes in hydraulic conductivity down core correlate with secondary dissolution and
cementation in the meteoric environment. In the upper reef crest sequence, cementation
and recrystallization increase with core depth. Vadose diagenesis is more prominent in
these sections where characteristic cements are matrix microspar and blocky calcite
inside moldic pores. As a result, interparticle porosity preservation is more abundant.
The lowest K values are in the reef front where thin-sections show fringe cements
occluding interparticle pores in the phreatic zone. Consequently, there are fewer
pathways for fluid flow. In addition, micritic matrix areas are also more common in the
reef front and usually have isolated moldic porosity with low pore connectivity. Core
recovery was used as a proxy for porosity. There is an inverse correlation with the
hydraulic conductivity. In the more massive and sandier zones where cementation is
more pe rvasive, core recovery is higher than the muddier and moldic branching coral
section of the back-reef and reef front.
It has been suggested that petrophysical data obtained from small-scale plugs may not
accurately reflect heterogeneities at a scale relevant to fluid flow within a reservoir. A
preliminary correlation between the permeability values calculated from plugs (Ditya,
2012), and the hydraulic conductivity calculated from in situ packer tests suggests that in
this example small-scale plug data can be reasonably scaled up for geomodels, however
exceptions do exist. The lack of correlation in the upper 5 m may just be a reflection of
the low resolution of permeability data or dissolution characteristics close to the
subaerial exposure not represented at the plug scale.
When completed, the study will provide a facies/diagenesis/acoustic-based correlation
of permeability (hydraulic conductivity) data after several stages of post-depositional
stabilization. These correlations can then be used as formation properties, and input as
model parameters as part of the upscaling for reservoir flow simulations.
REFERENCES
Ditya, A., 2012, Petrophysical characterization of Pliocene-Pleistocene reefal carbonates, southern
Dominican Republic: University of Miami M.S. Thesis, Open Access Theses, Paper 358,
http://scholarlyrepository.miami.edu/oa_theses/358.
Fogg G. E., and J. F. Lucia, 1990, Reservoir modeling of restricted platform carbonates:
geologic/geostatistical characterization of interwell-scale reservoir heterogeneity, Dune Field,
Crane County, Texas: University of Texas BEG Report Investigation 190, 65 p.
Fraser G. S., and J. M. Davis (Eds), 1998, Hydrogeologic models of sedimentary aquifers: SEPM
Concepts Hydrogeological Environmental Geology, v. 1, 188 p.
Lucia, J. F., 2007, Carbonate reservoir characterization: An integrated approach: Springer, 336 p.
20
ORIGIN AND DIAGENESIS OF MICROBIALITES ON THE
UPLIFTED ATOLL OF MARÉ, NEW CALEDONIA
Chelsea L. Pederson, Donald F. McNeill, James S. Klaus, and
Peter K. Swart
KEY FINDINGS
ƒ
Maré microbialites consist of aragonitic/HMC and intraclast nuclei surrounded
by a 0.1-3 mm microbial envelope
ƒ
Microbialite formation is interpreted to have formed just prior to subaerial
exposure, in shallow, agitated marine conditions
ƒ
Despite subaerial exposure of nearly 3 my and pervasive meteoric cementation,
the grain nuclei remain largely unaltered from their original structure and
mineralogy
ƒ
The constructive microbial envelope deters meteoric dissolution and formation of
secondary moldic porosity
SIGNIFICANCE
The influence of microbes on
carbonate textures is well known
from studies of stromatolites that
date back to the Precambrian. In the
Phanerozoic, with the diversification
of life forms, the dominance and
influence of microbes in the marine
realm has been largely relegated to
niche environments.
However,
microbial influence on carbonate
textures and recognition as distinct
facies
has
been
increasingly
recognized,
especially
in
the
calibration
of
geological
and
reservoir
models.
This
study
contributes to the understanding of
how
microbial
processes
can
influence rock textures with the Figure 1. Global Mapper topographic image of Maré, New
Caledonia.
characterization
of
proposed
cyanobacterial formed grains associated with the final stages of marine deposition of
Maré, New Caledonia (Figures 1 & 2) and the onset of uplift-driven subaerial exposure.
By characterizing the chemical and physical attributes of these microbial carbonates we
will provide a better understanding of their formation, environment of origin, and
influence on early diagenesis.
21
Figure 2. Cross section of the La Roche faro, where the inter-bedded microbialites occur.
RESULTS
Physical Characterization
Reefal units representing the final stages of marine deposition of the Maré atoll, New
Caledonia (Figure 1 & 2) are inter-bedded with horizontal units of large coated grains.
The coated grains range in size from 1-10mm. Each grain contains a nuclear fragment
surrounded by a microbial coating with irregular concentric layering (Figure 4a).
Meteoric cements bind the coated grains and partially fill the interparticle pore space.
The irregular concentric grains with diameters <10cm are therefore termed oncolites.
The general composition of the oncolite beds consists of a minimum of three distinct
components (Figure 4b). The innermost layer is the nuclei, made of aragonitic skeletal
fragments and intraclasts. Aragonitic mollusk shell fragments, some with their original
nacre (mother-of-pearl) are sometimes present. Nuclei thickness ranges from 0.604mm. The second layer is a microbial coating (rind) surrounding the nuclei. Rind
thickness ranges from 1.5-6 mm, representing up to 50% of the total oncoid diameter.
While most nuclei are not well rounded, the encircling rind generally displays a more
circular shape. The microbial rinds consist of fairly equant euhedral 1-3 micron diameter
crystals, tightly packed, with relatively low-permeability (Figure 4b & 4c). Point counting
has determined the microbialite samples to have porosities of less than 5%, with some
thin section samples having less than 1.5%. The third component of the oncolite grains
consists of blocky meteoric cement, representing 25-35% of the total sample area. The
coarse low-Mg calcite crystals encase the grains, yet the meteoric fluids appear to not
have penetrated the dense microbial rind, allowing the preservation of the inner nuclei.
Geochemical Characterization
The average į18O signal for the nuclei, rind, and cement layers were -4.7, -4.9, and -5.6
respectively. The average į12C signal for the nuclei, rind, and cement layers were -6.7, 7.6, and -10.1. These stable isotopic results show a general trend toward lighter carbon
and oxygen isotopes as you move from the nuclei to the cement (Figure 3). This trend
records diagenetic alteration of the microbial rinds by freshwater.
22
Figure 3. Stable isotopes of carbon and oxygen measured in ‰VPDB for the cement, rind, and nuclei
layers of Maré oncolites.
INTERPRETATION AND IMPLICATIONS
One key feature of the microbial rind is its influence on meteoric diagenesis. Unlike
subaerially exposed ooid grainstones, the microbial rind of the large oncolites inhibits
nuclei dissolution and the formation of associated moldic porosity. These lowpermeability rinds form an effective seal around the skeletal or composite nuclear grains.
The original mineralogy of the rind is unknown, but it was likely low-Mg calcite based on
crystal shape and absence of diagenetic features. Fibers in the rind, likely of microbial
origin, indicate in situ biological formation (Figure 4d). A layer of blocky calcite spar
cement, typical of meteoric conditions, encases the concentric oncolitic grains. The sharp
boundary between the coated grain and these calcite cements suggest distinct
environments of formation. The microbial rind promotes a rock fabric of well-preserved
aragonitic nuclei. Whether aragonitic or calcitic, the permeability of the microbial rinds
was sufficiently low at this early stage, to preclude freshwater access to the aragonitic
fragments in the nuclei. This exception to a typical diagenetic sequence is well illustrated
in thin section and SEM images and has important implications for the dissolution of
aragonitic grains during exposure. Furthermore, the recognition of microbial textures,
their formation, and distribution within existing facies models may be an exception to
the diagenetic paradigm.
REFERENCES CITED IN PRESENTATION
Guyomard, T., Aissaoui, D. M., and D. F. McNeill, 1996, Magnetostratigraphic dating of an
uplifted atoll, Maré Island, Loyalty Archipelago, S.W. Pacific: Journal of Geophysical
Research, v. 101, p. 601-612.
McNeill, D. F., and A. Pisera, 2010, Neogene lithofacies evolution on a small carbonate platform
in the Loyalty Basin, Maré, New Caledonienozoic carbonate systems: SEPM Special
Publication, v. 95, p. 243-255.
23
Figure 4. (A) Short core drilled June 2012 from the upper unit of the Maré atoll. Depth increases from top
left to bottom right. Reefal units comprise the upper and lower portions of the core, with a stratified
oncolite unit in the center. (B) Thin section photomicrograph of an oncolite sample from the top of an
uplifted isolated reef (faro) near the town of La Roche, Maré, New Caledonia. (C) SEM image of biofilm
remnants in the form of calcified cyanobacterial sheaths. (D) SEM image of low-permeability equant
calcite crystals in the microbial rind.
24
MICROBIAL COMMUNITY CHARACTERIZATION AND
FUNCTIONAL GENE DIVERSITY OF OOLITIC GRAINS FROM
GREAT BAHAMA BANK
Mara R. Diaz, Alan M. Piggot, Gregor P. Eberli, and James S. Klaus
KEY FINDINGS
ƒ
Confocal microscopy and measurements of extracellular polymeric substances
(EPS) on oolitic grains confirm the ubiquitous presence of attached biofilm
communities with variations related to hydrodynamic conditions.
ƒ
DNA sequencing reveals oolitic sediments harbor diverse bacterial communities
that differ between depositional environments (e.g., active ooid shoals, non active
ooid shoals, mat-stabilized calcareous sediments).
ƒ
Differences in microbial functional gene diversity reflect distinct biogeochemical
environments associated with active, non-active, and mat-stabilized sediments.
SIGNIFICANCE
Modern marine ooids are found in the Bahamas, the Yucatan, Arabian Gulf, South
Pacific, and Shark Bay, Australia. In these regions, ooid shoal complexes can form
margin parallel marine sand belts, tidal bar belts or platform interior blankets that
stretch up to 100 km. While it is widely recognized that ooid formation and early marine
cementation requires water supersaturated with respect to calcium carbonate, and an
agitated environment that allows for CO2 degassing, the role of microorganisms has been
debated for decades.
The physical presence or metabolic activity of microbes could influence the formation
or early cementation of ooids. Microbial metabolism in the form of photosynthesis,
anoxygenic photoautotrophy, sulfate reduction, denitrification, and ammonification can
promote calcification by creating a more alkaline environment, whereas aerobic
respiration, sulfide oxidation and ammonium oxidation can promote dissolution
(Dupraz and Visscher, 2005). In addition, production and degradation of EPS by
cyanobacteria and heterotrophs can either facilitate or inhibit CaCO3 precipitation. When
abundantly present, the EPS acts as a “sponge”, inhibiting carbonate precipitation by
trapping free divalent cations (i.e. Ca+2 and Mg+2). However, degradation of EPS by
heterotrophs (ie. sulfate reducers, denitrifiers, etc) reduces the cation binding capacity of
EPS by releasing free Ca+2, which triggers CaCO3 precipitation. The role of sulfate
reducer bacteria (SRB) in EPS production and degradation in the lithifying mats of the
stromatolites in the Bahamas has shown the importance of these microbial communities
and associated metabolisms, which are considered the environmental engine that drives
CaCO3 precipitation in lithifying microbial mats and stromatolites (Reid et al., 2000).
To better understand microbe-carbonate interactions we used a combined approach of
clone library sequencing, functional gene-based microarray (GeoChip 4) and confocal
laser scanning microscopy (CSLM) to study the microbial structure, functional gene
diversity and metabolic potential of microbial communities associated with oolitic grains
from active shoals, non-active and mat associated environments.
25
RESULTS
EPS determinations
Confocal image analysis of oolitic grains stained with cyanine die-conjugated lectin
demonstrated that all three environments harbored EPS-biofilm communities but their
densities were different as a clear progression was seen in the amount of EPS coating
from active to the mat-stabilized environment. These results are confirmed by phenolsulfuric acid quantification of ooid EPS levels, which on average showed mat-stabilized
EPS content to be ~10- 3.8 fold higher than active and non-active samples, respectively.
Figure 1. CLSM
images of ooids and
EPS measurements
derived from active,
non-active and
mat-stabilized
environments of
Joulters Cay,
Bahamas.
Phylogenetic Diversity
Bacterial communities were highly diverse and dominated by Proteobacteria (50-61%).
Difference in microbial composition among the three environments were attributed to
phylotypes within Proteobacteria and likely related to differences in the hydrodynamic
characteristics of each environment, which could affect the biogeochemistry of inorganic
nutrients and organic carbon availability. For instance, active ooid shoals are exposed to
severe hydrodynamic forces that could hamper deposition of sedimentary organic
matter.
Functional Metabolic Genes
To understand the diversity and functional capabilities of oolitic microbial
communities, we used a high-throughput functional gene microarray. Geochip 4 gene
array contains 83992 distinct probes covering 410 gene families associated with
microbial functional processes. This study recovered a total of 12,432 genes, representing
~14.8% of the total number of genes in the array. Phylogenetically, up to 39 different
lineages were documented representing 64 species from archaea, 823 from bacteria, 70
from fungi and 10 from others. The total number of genes varied significantly among the
environments. Active communities retrieved the least number (8631); followed by matstabilized (9455) and non-active communities (10,102).
Carbon fixation. All three environments were characterized by high levels of carbon
fixation genes associated with the Calvin–Benson CO2 pathway (RuBisCO). While active
environments recovered the lowest signal intensities, non-active and mat environments
recovered higher signals (Figure 2). Other autotrophic CO2 fixation processes included
carbon monoxide dehydrogenase (CODH), BchY and ATP citrate lyase genes (aclBATP). The most significant differences in the level of CODH, was observed between
active and non-active communities. In addition, all three environments recovered aclBATP and BchY genes, both of which are employed as biomarkers for anoxygenic
phototrophs (Figure 2). BchY genes, were found to exhibit significantly higher levels in
non-active environments (non-active vs mat: P<0.001; non-active vs active: P<0.001).
26
Figure 2. Distribution of
key genes involved in
carbon cycling processes.
The signal intensities were
the
sum
of
detected
individual gene sequences
for each functional gene.
Gene classifications are as
follows: Rubisco, CODH,
aclB and BchY (autrophic
Co2 fixation); mmox and
pmoA (methane oxidation);
mcrA
(methane
production) and FTHFS
(acetogenesis).
Acetogenesis and Methane Metabolism. Methane production (mcrA), methane
oxidation (mmox, pmoA) and acetogenesis (FthFS), do not appear to play a key role in
oolitic environments since genes associated with the aforementioned metabolisms were
~5-12 fold lower than other autotrophic CO2 fixation genes e.g. RuBisCO, CODH genes
(Figure 2).
Carbon degradation. Carbon degradation appears to be important in oolitic
environments as genes involved in carbon degradation were highly enriched. Carbon
degradation genes were mostly affiliated with Bacteria, representing 65.3% of the total
pool of C degrading genes, whereas fungi (e.g. Ascomycota and Basidiomycota)
contributed 33.4%. Other minor contributors included members within the Archaea
phyla e.g. Euryarchaeota and Crenarchaeota. The most labile forms of carbon, e.g.
starch, hemicellulose and cellulose, recovered a wider array of enzymes (except for van
A, acetylglucosaminidase, exochitinase, and mnP) than recalcitrant compounds e.g.
aromatics, chitin, lignin and pectin.
Nitrogen Cycling. Five enzymes involved in denitrification including nitrate reductase
(narG), nitrite reductase (nirK, nirS), nitric oxide reductase (norB) and nitrous oxide
reductase (nosZ) were documented. When pooling all five functional genes, signal
intensities were significantly different among sites (P= 0.01 to 0.001), with higher levels
documented for non-active and mat-stabilized environments. Besides genes involved in
denitrification, Geochip detected two key enzymes involved in ammonification processes
e.g. glutamate dehydrogenase (gdh) and urease c (ureC), both of which showed
significant differences (P=0.02 to 0.001) among sites.
Sulfur cycling genes. Most functional genes involved in sulfur cycling were detected.
Genes related to sulfite reductase (dsrA/B, cysJ/I, SiR) and sulfur oxidation (e.g. soX)
were most abundant, whereas genes related to sulfide oxidation (e.g. sqR, fccAB), and
dissimilatory adenylsulfate reductase were less abundant (e.g. AprA/B). As previously
documented, active environments consistently displayed lower signal intensities as
opposed to mat and non-active environments. When combining all sulfite reduction
genes, significant differences between pairwise comparisons of active vs mat (P = 0.001)
and active vs non-active (P = 0.009) were detected. Sulfur oxidation recovered a total of
65 sox genes, most of which were associated with Alphaproteobacteia and
Gammaproteobacteria.
27
INTERPRETATION AND IMPLICATIONS
Differences in the bacterial communities of active, non-active, and mat-stabilized
environments could influence nutrient recycling through variations in the level of both
primary production and remineralization processes. Bacterial primary production
appears to be driven by a diverse population of autotrophic CO2 fixers among which,
Proteobacteria (Alpha-Beta, Gamma), Chlorobi, and Cyanobacteria are the most
prevalent. This composite of microbes is not only aerobic phototrophs but includes
anaerobic/aerobic anoxygenic phototrophs. All of these can drive the alkalinity towards
carbonate precipitation by consuming bicarbonate and increasing local pH.
Carbon degradation appears as a major biological process in oolitic sediments. The
widespread array of C substrates suggests these microbial communities have a high
plasticity in metabolic pathways that enables them to degrade complex organic matter
and exploit limited carbon sources, which are commonly associated with oligotrophic
waters. For instance, the potential ability of microbes to use chitin, cellulose, and lignin
as major carbon, nitrogen and energy sources has been documented in marine
environments (Sieburth, 1976).
The presence of denitrifiers and sulfate reducers in active ooid sediments suggest their
metabolic pathways are not constrained to non-active and mat-stabilized environments.
Their ubiquitous occurrence is in agreement with earlier studies that documented their
prevalence in sediments of the Bahama Bank (McCullum, 1970). In Drew (1911)
carbonate precipitation in mud flats (whitings) appears to be mostly driven by the action
of denitrifiers on the calcium salts present in seawater.
We also document and measure the abundance levels of EPS on ooid grains. Both,
confocal microscopy and phenol-sulfuric acid methods showed increasing amounts of
EPS from active to mat stabilized environments. While the synergistic effects of EPS and
heterotrophs (i.e. SRB) on micrite laminae and lithification in microbialite systems have
provided important insights into the inteplay of microbial communities and chemical
processes that regulate CaCO3 precipitation, other studies have focused on the role of
EPS and associated photosynthetic communities in the formation of the carbonate cortex
in freshwater ooids. The waters of the Bahamas Archipelago are supersaturated with
respect to CaCO3 and ideal for precipitation. However, based on pervasive EPS as well as
microbes with the potential to influence carbonate precipitation, abiotic and biological
factors are probably intertwined in the precipitation processes that form ooids, marine
cements, and carbonate mud.
REFERENCES
Dupraz C., and P. T. Visscher, 2005, Microbial lithification in marine stromatolites and
hypersaline mats: Trends in Microbiology, v. 13, p. 429-438.
Drew, G. H., 1911, The action of some denitrifying bacteria in tropical and temperate seas and the
bacterial precipitation of calcium carbonate in the sea: Journal of the Marine Biological
Association, v. 9, p. 142.
McCallum, M. F., 1970, Aerobic bacterial flora of the Bahama Bank: Journal of Applied
Bacteriology, v. 33, p. 533-542.
Reid, R. P., Visscher, P. T., Decho, A. W., Stolz, J. F., Bebout, B. M., Dupraz, C., Macintyre, I. G.,
Paerl, H. W., Pinckney, J. L., Pufert-Bebout, L., Steppe, T. F., and D. J. DesMarais, 2000, The
role of microbes in accretion, lamination and early lithification of modern marine
stromatolites: Nature, v. 406 p. 989-992.
Sieburth, 1976, Bacterial substrates and productivity in marine ecosystems: Annual Review of
Ecology, Evolution, and Systematics: v. 7, p. 259-285.
28
PORE STRUCTURE AND PETROPHYSICAL
CHARACTERIZATION OF MICROBIALITES
Gregor P. Eberli, Ralf J. Weger, Jan Norbisrath, and
Giovanna della Porta1
1)
University of Milano
KEY FINDINGS
ƒ
Microbialites such as stromatolites, travertine and microbially cemented
hardgrounds have simple pore structures of highly variable size.
ƒ
The pore structure of microbialites produces a stiff frame that results in:
ƒ
o
high velocity at a given porosity
o
good pressure resistance
o
maintenance of primary porosity to great burial depth
Resistivity is high compared to other carbonates as a result of scattered and
isolated pores.
SIGNIFICANCE
Microbialites are one of the major reservoir facies in the pre-salt offshore Brazil and
stromatolites are an important reservoir facies in some Proterozoic carbonate reservoirs.
The reason why the microbialites maintain good reservoir quality to a great burial depth
is their remarkable amount of preserved primary porosity.
This study investigates the porosity, permeability, sonic velocity and resistivity, in
conjunction with the pore structure, to identify the petrophysical characteristics of
microbialites. The data set and its interpretation are intended to help explain the
petrophysical signature of microbialites in log and seismic data. The incorporation of
modern microbialites helps to capture the microbial processes that produce the
characteristic petrophysical properties of microbialites.
Figure 1. The data set for this study consists of microbially cemented submarine hardgrounds from the
southern end of the Tongue of the Ocean (left), modern stromatolites from the Bahamas (middle), and 70
travertine samples from unknown quarries in Italy (inset = photomicrograph displaying travertine
porosity). The travertine samples were collected and classified by Giovanna della Porta and her team.
29
INFLUENCE OF MICROBIAL ACTIVITY ON PETROPHYSICAL PROPERTIES
Microbial activity influences petrophysical properties because it 1) promotes
mineralization and dissolution, 2) can fuse grains together, 3) can precipitate microbial
boundstone, and 4) produces unique pore structures. Most of these processes positively
affect the rock stiffness. As a result, microbialites are petrophysically characterized by
high velocity, high resistivity and relatively high porosity. The early microbial processes
that construct and strengthen the rock are also responsible for their ability to resist
compaction. This early stiffening of microbialites is seen in modern hardgrounds and
stromatolites where velocity does not, or only slightly increases with pressure (Figure 2).
Likewise, travertine samples display a small but consistent increase of velocity with
increasing confining pressure.
Figure 2.
Velocity evolution of
modern hardgrounds
(TOTO), stromatolites
and travertine samples
with increasing
pressure.
Most hardgrounds and
stromatolites do not
display a velocity
increase with increasing
confining pressure.
Travertines have a high
acoustic velocity and
display a small increase
with increasing
pressure.
VELOCITY AND PORE STRUCTURE OF MICROBIALITES
Table 1 provides an overview of the measured petrophysical properties. Porosity is
lower in the travertine samples compared to the marine hardgrounds and stromatolites.
Microbialites have a high velocity despite their high porosity (Figure 3). In particular,
stromatolites and travertine samples display high velocities even at high porosity.
Table 1. Petrophysical Properties
Ø
Travertine
0.4 – 23.4
Hardground
15 – 41.6
Stromatolite
12.8 – 30.1
30
K (mD)
Vp (dry)
Vs (dry)
Vp (wet)
Vs (wet)
0.01 – 1470
4658 – 5976
2611 – 3233
4891 - 6131
2297 - 3183
3266 - 4986
1513 - 2662
4407 - 5402
2334 - 2999
70 - 1694
m
1.9 – 5.8
2.1 – 2.6
The high velocities of microbialites are caused by their pore structures. The pore types
are mostly interparticle and intraframe (only observed in the travertines). In rock
physics models intergranular porosity rocks are usually considered to have compliant
pores and low stiffness, and therefore low acoustic velocity. In contrast, rocks with
moldic and vuggy porosity are classified as stiff, high aspect ratio rocks, resulting in high
bulk modulus and velocity (Lucia and Conti, 1987). Microbialites have a high velocity
because microbial processes weld grains together in a form of micritic bridging cement
or micritic crusts to form a strong frame of intergranular porosity (Hillgärtner et al.,
2001). Despite the microbial bindings the pore structure is relatively simple, resulting in
a low Perimeter over Area (PoA) when quantified with digital image analyses. Yet, the
pore size varies considerably. Rocks with large simple pores are generally fast (Weger et
al., 2009). The measured microbialites follow this trend.
Figure 3.
Velocity -porosity plot of
the three data sets and a
comparison of the data
set of Weger et al.
(2009).
In samples with
overlapping porosity
travertine has the
highest acoustic velocity
at a given porosity,
followed by
stromatolites. Marine
hardground samples
(TOTO) have the highest
porosity and high
velocity.
Figure 4.
Characterization of the
pore system in
microbialites with digital
image analysis
parameters compared to
the data set of Weger et
al. (2009) and Verwer et
al. (2010)
The microbialites have
simple (small PoA) pores
of variable size
(DOMsize).
31
RESISTIVITY
Resistivity as expressed by the Formation Factor is high in travertines. The
cementation factor “m” in the travertine samples displays a very wide range from 1.9 –
5.8. In comparison, “m” is 1.72 – 4.14 in the data set of Verwer et al. (2011) that included
a variety of carbonate textures and pore types.
The high resistivity of the travertines has its roots in the unconnected pore structure
and the relatively small number of pores. Macropore structures are simple and variably
sized (Figure 4). These pores, like many of the micropores, are very scattered and porous
areas are isolated by very dense areas without visible porosity. This, together with the
low pore count, results in the high cementation factor and high resistivity.
Figure 5.
Formation factor vs.
porosity of the travertine
compared with
stromatolites and the
data set of Verwer et al.
(2011).
REFERENCES
Hillgärtner, H., Dupraz, C. and W. Hug, 2001, Microbially induced cementation of carbonate
sands: are micritic cements indicators of vadose diagenesis? Sedimentology, v. 48, p. 117-131.
Lucia, F. J. and R. D. Conti, 1987, Rock fabric, permeability, and log relationships in an upwardshoaling, vuggy carbonate sequence: The University of Texas at Austin, Bureau of Economic
Geology, Geological Circular 87-5, 22 p.
Weger R. J., Eberli, G. P., Baechle, G. T., Massaferro, J. L., and Y. F. Sun, 2009, Quantification of
pore structure and its effect on sonic velocity and permeability in carbonates: AAPG Bulletin,
v. 93, no. 10, p. 1-21.
Verwer, K., Eberli, G. P., Baechle, G. T., and R. J. Weger, 2010, Effect of carbonate pore structure
on dynamic shear moduli: Geophysics, v. 75, p. 1–8.
Verwer, K., Eberli, G. P. and R. J. Weger, 2011, Effect of pore structure on electrical resistivity in
carbonates: AAPG Bulletin, v. 95, p. 175-190.
32
ROCK FLUID INTERACTION: HOW DISSOLUTION
INDUCED CHANGES IN PORE STRUCTURE AFFECT
ACOUSTIC VELOCITY
Ralf J. Weger, Peter K. Swart, Gregor P. Eberli, and Mark Knackstedt
KEY FINDINGS
ƒ
Carbonate rocks don’t need millennia to change with respect to acoustic
properties.
ƒ
Observed changes in velocity during dissolution are significantly smaller than
expected from porosity increases and predicted by rock physics models.
ƒ
Dissolution occurs preferentially on sub-micron scales.
ƒ
Pore structure simplification during dissolution inhibits drastic decreases of
acoustic velocity.
SIGNIFICANCE
A large number of publications examine the influence of temporary, reversible effects of
pore fluids on the rocks framework and its acoustic properties (e.g. Baechle et. al., 2009).
However, assessments of the influence of cementation and dissolution on the rock
stiffness, porosity preservation and acoustic properties are rare. This study aims to shed
light on how carbonate reservoir rocks behave in contact with formation fluids.
Laboratory experiments on precipitation and dissolution of carbonates confirm that the
chemical reaction with the pore fluid is a continuous process that changes both fabric
and pore structure within days. The changes caused by dissolution and/or precipitation
include alteration of the rocks pore structure, porosity, permeability, and its acoustic
properties. As a result, the comparison of different vintages of acoustic data either from
logs or from time-laps seismic data carries high uncertainties.
Figure 1. High-resolution CT-scan of a rudist grainstone before (left) and after (right) the dissolution
experiment. The light color is the solid portion, the gray is the pore space. An increase of the grey color
from the left to right documents the increase of pore space by ~10%.
33
This study examines and attempts to quantify the changes that occur during
precipitation and dissolution in a controlled laboratory setting, specifically those
affecting porosity, permeability and sonic velocity in carbonates. We observe both an
increase in rock stiffness in the transition from sediment to rock during precipitation,
and drastic alterations of the rocks fabric and pore structure during dissolution. CTscans before and after dissolution are used to correlate the observed changes in physical
properties to the changes in the rock fabric and also to determine where exactly the
dissolution took place and what portions of the rock were altered (Figure 1).
EXPERIMENTAL SETUP
The experimental setup allows for precipitation or dissolution of calcite cement by
filtration of pore fluids and the simultaneous measurement of acoustic velocity. A
filtration system provides continuous flow of fluids that are supersaturated (or undersaturated) with respect to CaCO3 through the rock samples and simultaneously monitors
the fluid properties. Any changes are immediately compensated for by altering the fluid
to maintain the original fluid composition. An Aqua Medic SP 3000 peristaltic pump
provides a continuous flow rate of 50 ml/min or less. Temperature and pH of the pore
fluid are monitored and logged in 5 sec intervals.
Although precipitation of calcite cement is relatively fast, it takes several days-weeks to
produce a rock from loose sediment grains. Initially the sediment is compacted to 5 MPa
and subsequently saturated. A stepwise initial pressurization to 40 MPa ensures full
compaction. Throughout the experiment the confining pressure is held at 10 MPa and
the pore pressure is given by the flow rate. Acoustic velocity is measured in one-hour
intervals. After the experiment samples are analyzed under SEM to determine where
and how many crystals have formed during the experiment. Several precipitation
experiments were performed during which the CaCO3 concentration was kept close to
but above saturation for ~90 hrs.
Dissolution occurs much faster than precipitation; initial experiments were performed
for 5 days, and a subsequent dissolution experiment was performed for 3 days while
maintaining the pore fluid at a pH no greater than six. Both SEM images and CT –scans
where acquired before and after the experiment to visualize and analyze the changes
(Figure 1). Particular attention was paid to determining where the dissolution of calcium
carbonate had occured. CT-scan data was calibrated and segmented into pore space,
rock space, and intermediate (microporous) space (Figure 2).
PRECIPITATION EXPERIMENTS
All experiments produced permanent alteration of the rocks within only hours or days.
The precipitation experiments show small but measurable increases in acoustic velocity
and decrease of porosity. SEM analysis revealed variable amounts of small crystals
precipitated on the ooids. Two distinct types of crystallization are identified: 1) crystals
formed directly at the grain to grain contact, fusing the ooids together into a more
coherent unit, causing an increase in acoustic velocity of ~50 m/s to ~300 m/s; 2)
precipitation occurred in the form of needle-like crystals.
In samples where precipitation creates a stiffer framework by fusing grains together the
observed increase in velocity is comparable with model prediction (Extended Biot
Theory model) and can be attributed to the fusing of grain-grain contacts. In other
samples, velocity increased throughout the experiment from ~2200 m/s to ~2500 m/s
(3400 m/s to 3700 m/s), but the change is disproportionally small with respect to the
34
Figure 2. Comparison of high-resolution CT-scan slices before and after the experiment. On these slices
pore space is black. The difference in pore space before and after the experiment is given in grey tones in
the picture to the right. This provide the visualization of the porosity enhancment and illustrates that
alterations are expanding upon existing porespaces and pathways, and decrease the amount of
microporosity.
porosity decrease from over 36% (35%) to ~25% (28%) that occurred during the
experiment. This small velocity increase is attributed to the needle-like crystals that
formed in these samples, which do not increase the stiffness of the rock but produce
small particles in the inter-grain space and create a more complex pore-system.
DISSOLUTION EXPERIMENTS
Several dissolution experiments were performed during which the CaCO3 concentration
was kept close to but below saturation for 3-5 days. Pleistocene ooid grainstone and
Cretaceous rudist grainstone samples were used, and both SEM images and CT–scans
were acquired to analyze the changes. During all experiments porosity increased
drastically by 10-15%. The dissolution experiments show small decreases in acoustic
velocities from ~3200 m/s to ~3000 m/s.
In the first two samples, channel-like dissolution is observed and appears to leave the
stiff part of the rock frame intact while substantially increasing the rocks porosity. Small
fragile particles and small grains appear to be dissolved first and the proportion of
microporosity decreases. CT-scan analysis shows that the proportion of microporosity
drastically decreases during the experiment (Figure 3).
Measuring the Perimeter over Area (PoA) and Dominant (DOM) size using digital
image analysis indicates that the process of dissolution simplified the overall pore
geometry (Figure 4). CT-scan derived parameters show clear increases in DOM size and
decreases in PoA.
Figure 3. Comparison of
macro
and
micro
porosity derived from
individual slices of CTscans before (blue) and
after (red) dissolution
experiment.
After
dissolution experiment,
nearly all micro porosity
has disappeared.
35
The large increase in porosity combined with the simplification of the pore geometry
due to selective dissolution make the frame flexibility substantially lower after
dissolution than it was before. The lower frame flexibility and the less complex pore
geometry result in only moderate velocity decrease during dissolution.
Figure 4. (Bottom) Velocity-porosity cross plots of a diverse set of samples (Weger et al., AAPG 2009) with
theoretical values for frame flexibility in the background. Color represents geometrical parameters
Perimeters over Area (PoA) and Dominant Size (DomSize). (Top left) Thin section images corresponding to
the nine enlarged dots on the cross-plots. (top right) Digital image parameters PoA and DOMsize obtained
from CT-scans of dissolution sample before and after dissolution. (Inside red eclipse) Dissolution sample
before and after dissolution with DIA parameters in color.
REFERENCES
Baechle, G. T., Eberli, G. P., Weger R. J., and J. L. Massaferro, 2009, Changes in dynamic shear
moduli of carbonate rocks with fluid substitution: Geophysics, v. 74, no. 3, p. 135-147
Weger R. J., Eberli, G. P., Baechle, G. T., Massaferro, J. L., and Y. F. Sun, 2009, Quantification of
pore structure and its effect on sonic velocity and permeability in carbonates: AAPG Bulletin,
v. 93, no. 10, p. 1-21.
36
NEW INSIGHTS INTO SLOPE PROCESSES FROM THE
BAHAMAS AND WEST FLORIDA
Gregor P. Eberli, Donald F. McNeill, Thierry Mulder1,
Emanuelle Ducassou1, Dierk Hebblen2, Claudia Wienberg2, and
Paul Wintersteller2
1) University of Bordeaux)
2) University of Bremen
KEY FINDINGS
ƒ
New data from the north slope of Little Bahama (LBB), the western Great
Bahama Bank (GBB) and the west Florida shelf refine existing models of
processes, morphology and facies distribution along carbonate slopes.
ƒ
The steep platform margin is less prone to failure than previously thought. In
contrast, large-scale slope failures occur on the low-angle mid–lower slopes that
produce extensive carbonate mass transport complexes (MTC) on the toe-ofslope.
ƒ
The slope is dissected by a variety of incisions that include regularly spaced
gullies (20 – 30 m deep), a series of channels (up to 60 m deep), and spectacular
slope canyons (up to 200 m). Channelized and canyon-dissected areas abruptly
end laterally.
ƒ
Creeping of the slope is common on many places along strike.
ƒ
The size and abundance of the debris of the MTC is highly variable but where
present the debris serves as a foundation for an extensive deep-water ecosystem.
SIGNIFICANCE
Modern carbonate platforms occur in a variety of settings as attached or isolated
platforms and may be surrounded by steep slopes that can exceed 45º. Early work in the
Bahamas contributed to the understanding of the slope anatomy and facies distribution,
in particular carbonate turbidites (Mullins et al., 1984). One of the objectives of the first
leg of the Ocean Drilling Program (ODP) was to explain the slope variability of
accretionary, bypass and erosional slopes (Austin et al., 1988). The evolution of the
prograding platform and slope within a sequence stratigraphic context was documented
by seismic and coring during ODP Leg 166 (Eberli, 2000). Subsequent submersible
expeditions added information about the role of early diagenesis on slope stability
(Grammer et al., 1993) and carbonate redeposition in respect to sea level and ocean
currents (Anselmetti et al., 2000). Several of these studies have made it into textbooks as
modern analogs for carbonate slopes (Playton et al., 2012).
Yet, what was still missing was a comprehensive overview of the seafloor morphology
that would delineate the dimension and relationships of the different facies elements of
the margin and slope. New multibeam and seismic data, combined with piston coring
and visual observation with a Remote Underwater Vehicle, that were collected in two
recent cruises in the Bahamas and Florida provide for the first time high-resolution
digital elevation maps of the slope on LBB and GBB and the edge of the West Florida
shelf.
37
Taken together these new data refine existing models of processes, morphology and
facies distribution along carbonate slopes. In particular, they indicate the importance of
slope instability on the lower slope, the lateral variability of slope canyons and the extent
of debris fields on the toe-of slope that serves as the habitat of a diverse deep-water
ecosystem.
DATA SETS
Figure 1. The data sets for this study are of two generations. In the eighties submersible dives at Chub Cay
and the Tongue of the Ocean investigated the platform margin and uppermost slope. Multichannel seismic
data along the western margin of GBB were collected in the nineties. The new generation of data are
collected in the last 10 years from a multitude of platforms: ship based multibeam data for the slope of
LBB in 2002; AUV- based high-resolution multibeam; side scan sonar and sub-bottom profiles at five sites
for cold-water coral environments in 2005; multibeam, single channel and multichannel seismic data and
piston cores were collected onboard R/V Le Suroit in 2010 north of LBB and west of GBB during the
Carambar cruise, chief scientists on board were Thierry Mulder and Emanuelle Ducassou of the University
of Bordeaux; multibeam acoustic and single channel seismic data, grab samples and piston cores, as well
as visual observations with an ROV revisited two sites at the western side of GBB and also on the western
edge of the West Florida shelf (not shown on map) in the spring of 2012 on board R/V Maria Merian
during cruise MSM 20-4 with Dierk Heblen as “Fahrtleiter”. The Carambar cruise was financed by the
French National Science Foundation and cruise MSM 20-4 was paid for by the German Funding agency.
THE PLATFORM MARGIN AND THE ONLAPPING WEDGE
On most modern carbonate platforms the platform edge is between 100 m to a few
kilometers seaward in water depths of around 20 – 30 m. Seaward of the platform edge,
a near vertical cemented slope of approximately 100 m height, called either the wall or
38
escarpment, develops on all Bahamian carbonate platforms on both the windward and
the leeward side (Grammer et al., 1993). The steep slope (45Û - 70Û) beneath the
escarpment is also cemented with a thin veneer of sediment and occasionally large
boulders and talus debris. For example along the southern side of LBB, only 2 - 28
boulders per kilometer are observed on multibeam data. No large-scale margin collapse
is observed on either LBB or on the northernmost 180 km along western GBB.
The steep uppermost slope beneath the escarpment is onlapped by soft sediment that
can reach up to 125 m in thickness in the Holocene. The wedge often displays a moat at
the point of contact indicating strong currents along the upper slope. The wedge thins
downslope and interfingers with coarse-grained sediment and debris from mass wasting.
THE SLOPE
The slope of western GBB where declivity is less than 4Û contains five major
morphological elements (Figure 2). They are 1) regularly-spaced gullies in the soft
sediment of the onlapping wedge, 2) numerous small and a few large-scale slope failures
in the slope of less than 2Ûthat produce scars up to 100 m in height and several
kilometers in length, 3) kilometer long scars that dissect the soft sediment, indicating the
continuous movement on the low angle slope. This scar is associated with features that
are interpreted as creeping of the slope. Single-channel seismic data and multibeam data
provide evidence that creeping is major slope process that is occurring along the entire
slope (Figures 3 and 4). Creeping seems to be restricted to the soft sediment portion of
the slope because it decreases as the sediment thickness of the mud wedge decreases. At
the toe-of-slope, the mud wedge thins and as a result older boulders and blocks from
earlier mass transport events litter the sea floor (Figure 4).
Figure 2. Multibeam bathymetry of approximately 50 km of the slope of western GBB, displaying four
major slope features. (1) Soft sediment wedge that is dissected by regularly spaced gullies, (2) large scars
and mass transport complexes, (3) long scars that displace the soft sediment wedge, and (4) partly buried
debris field on the toe-of-slope that documents earlier slope failure and mass transport debris.
39
Figure 3. (Left) Multibeam bathymetry of western GBB with location of close-up. (Right) Close-up of slope
above the scar of the MTC. The muddy slope shows an incipient slope failure with buckled sediment below
that are interpreted as the result of slope creeping.
Figure 4. (Left) Single-channel seismic profile taken during the MSM 20-4 cruise across a slope scar and
the modern sediment. The scar is approximately 50 m high. The most recent sediment is imaged as a thinly
layered succession. Down slope of the scar, these sediments form a wrinkled sea-bottom morphology due
to sliding along a surface (red in Figure). (Right) Multibeam bathymetry of the slope beneath the scar.
Divergent ridges that are dotted with boulders dominate the seascape; they are the debris field of an
earlier partly buried MTC .
40
Figure 5. Detail of slope canyons on the north slope of LBB (above) and a multichannel seismic section
across these canyons. The 3-D view of three canyons with coalescing arcuate scars (As) forming the
amphitheatre envelope, retrogressive erosion (Re) and terraces (T), talweg incision (Th) and channel (Ch).
Canyon numbering in red. Bottom) High-resolution multichannel seismic profile showing talweg incision
through morphologic terraces (from Mulder et al., 2012b)
Slope failures in the form of mass transport complexes (MTC) occur at several locations
on the lower slope The largest of these mass transport events consists of three scarps of
over 9 km length (Figure 2). The scar height ranges from 80 to 110 m (Figure 3). The
entire MTC covers an area of about 300 km2 (Mulder et al., 2012a). The slope angle
where this MTC is located is less than 2Û. At the west Florida shelf the entire shelf
breaks off at this low angle, indicating that major mass wasting events in carbonates do
not require a steep slope.
ROV observations, grab sampling and coring in both the channelized slope and the
MTC of GBB and West Florida document that the mass wasting debris is the foundation
for cold-water corals mounds. This colonization of antecedent topography explains why
neither mound shape nor orientation of the coral mounds along western GBB correlate
to the local current pattern (Correa et al., 2011). ROV images show boulders either partly
or completely covered with a diverse cold-water coral fauna (Figure 6).
41
Figure 6. (Left) Boulder partly colonized by deep-water fauna with surrounding coarse sediment. Water
depth 720 m; location west of Bimini (above). (Right) Block covered by cold-water corals and associated
fauna. Water depth 650 m, western margin of GBB. Photos taken with ROV during cruise MSM 20-4,
2012.
REFERENCES
Anselmetti, F. S. Eberli, G. P., and Z.-D. Ding, 2000, From the Great Bahama Bank into the
Straits of Florida: A margin architecture controlled by sea level fluctuations and ocean
currents: Geological Society of America Bulletin, v. 112, p. 829-846.
Austin, J. A., Jr., Schlager, W., et al., 1988, Leg 101—an overview: Proceedings ODP, Scientific
Results, v. 101, p. 455-472.
Correa, T. B. S., Grasmueck, M., Eberli, G. P., Reed, J., Verwer, K., and S. Purkis, 2012,Variability
of cold-water coral mounds in a high sediment input and tidal current regime, Straits of
Florida. Sedimentology, v. 59, p. 1278-1304, doi: 10.1111/j.1365-3091.2011.01306.x
Eberli, G. P., 2000, The record of Neogene sea-level changes in the prograding carbonates along
the Bahamas Transect—Leg 166 synthesis: Proc. ODP, Sci. Results, 166: College Station, TX
(Ocean Drilling Program), v. 166, p. 167–177.
Grammer, G. M., Ginsburg, R. N., and P. M. Harris, 1993, Timing of deposition, diagenesis, and
failure of steep carbonate slopes in response to a high-amplitude/high-frequency fluctuation
in sea level, Tongue of the Ocean, Bahamas: AAPG Memoir 57, p. 107-131.
Mulder, T., Ducassou, E., Eberli, G. P., Hanquiez, V., Gonthier, E., Kindler, P., Fournier, F.,
Leonifde, P., Billeaud, I., Marsset, B., E., Reijmer, J. J. G., Bondu, C., Joussiaume, R.,
Pakiades, M., 2012a, Morphology and sedimentary processes along a carbonate slope.
Example of the Great Bahama Bank: Geology, v. 40, p. 603-606, doi:10.1130/G32972.1
Mulder, T., Ducassou, E., Gillet, H., Hanquiez, V., Tournadour, E., Combes, J., Eberli, G. P.,
Kindler, P., Gonthier, E., Conesa, G., Robin, C., Sianipar, R., Reijmer, J. J. G., and A. François,
2012b, Canyon morphology on a modern carbonate slope of the Bahamas: evidence of a
regional tectonic tilting: Geology, v. 40, p. 771-774.
Mullins, H. T., Heath, K. C., Van Buren, H. M., and C. R. Newton, 1984, Anatomy of a modern
open-ocean carbonate slope: northern Little Bahama Bank: Sedimentology, v. 31, p. 141-168.
Playton, T. E., Janson, X., and C. Kerans, 2012, Carbonate slopes. In: Walker, R. G., and N. P.
James (Eds.), 2012, Facies model, response to sea level change: Geological Association of
Canada Press, p. 447-474.
42
VARIABILITY OF SLOPE AND BASIN FLOOR MORPHOLOGY
ALONG SOUTHWESTERN GREAT BAHAMA BANK
Andrew Jo, Gregor P. Eberli, and Mark Grasmueck
KEY FINDINGS
ƒ
Newly acquired multibeam bathymetry, backscatter, and sub-bottom profile data
provide magnificent visualization of sea bottom morphology that is highly
variable along strike.
ƒ
The steep margin is onlapped by an up to 125 m thick mud wedge that forms a
continuous moat of up to 30 m depth along the margin.
ƒ
A multiphase mass transport complex with an 11 km long slump scar in the upper
slope sheds large blocks (30 m high, 500 m length) 13 km into the basin.
ƒ
Backscatter and sub-bottom profile data reveal an unexpected but consistent
sediment distribution along the slope. The mud wedge forms the muddy ~ 6- 7
km wide upper slope, the middle to lower slope is ~ 20 km wide with coarsegrained deposits supporting channels and boulders that finally transition into the
fine-grained basin floor.
SIGNIFICANCE
Carbonate slope and basin floor reservoirs are considered underdeveloped in
hydrocarbon exploration. Sedimentary processes on carbonate slopes vary greatly
depending on various external and internal controls and hence such deposits are very
heterogeneous in composition, architecture, and lateral continuity (Playton et al., 2010).
This data set allows the heterogeneity of slope and basin floor morphology to be
documented along dip and strike. The over 100 km long data set is unique in assessing
the distribution, dimensions and variability of the slope facies that, together with ground
truthed data, will improve our understanding of slope to basin floor depositional
processes.
DATA SET
The data for this study was acquired by the Bahamas Petroleum Company. The study
area is located at the confluence of the Old Bahama and the Santaren Channel and covers
the western slope of Great Bahama Bank (GBB), the basin floor and northeast dipping
seafloor covering the SE-NW trending Cuban fold and thrust belt. The data include high
resolution multibeam bathymetry, backscatter, and single channel seismic survey in an
area of 6,512 km2, 742 lines totaling 5,342 km. The data provide unprecedented
visualization of slope and basin floor morphology along southwestern GBB (Figure 1).
43
THE DEPOSITIONAL ENVIRONMENTS OF SOUTHWESTERN GBB
Three main depositional areas are recognized in the study area (Figure 1).
1) The GBB margin and slope has slope angles that vary from 4Û - 76Ûand is
approximately 25 km wide. The main morphologic elements of the slope are: a steep
margin with an onlapping mud wedge, channelized slope with blocks and lobes of
redeposited carbonates, and mass transport complexes.
2) The basin floor is about 15 km wide and has homogeneous sediment cover. In some
places boulders and distal portions of the lobe reach the basin floor.
3) A northeast sloping seafloor with a slope angle of 0.08Û in the southwestern portion of
the study area is characterized by pockmarks in soft sediment.
Figure 1. Location of study area in the southwestern Great Bahama Bank and three different
depositional environments.
Upper Slope Morphology: margin and onlapping mud wedge
The morphology of the margin and upper slope varies from the northern to southern
portion of the study area. In the northern and mid portion, the margin has a declivity of
up to 30° (Figure 2.a), while in the southernmost portion of the study site, it is up to 76°.
A nearly transparent mud wedge onlaps the margin and thins basinward. At its
maximum extent it reaches 125 m in thickness and extends 4.2 km from the margin
(Figure 2.b). Outside of the study area where the wedge was cored it consists of 90%
fine-grained aragonite mud and is dated as less than 11 kyrs in age (Wilber et al., 1990;
Malone et al., 2001). This mud wedge is mostly off-bank transported lime mud. In the
study area the wedge is not just onlapping but it forms a moat that is 20- 30 m deep and
runs for tens of kilometers (Figure 2.b).
44
a
b
Figure 2. (a) Multibeam bathymetry from -50 to -350 m along Great Bahama Bank, displaying the steep
margin and the onlapping mud wedge with the characteristic moat. (b) Sub-bottom profile across the thick
mud wedge onlapping the margin. The wedge thins basinward across buried blocks.
Middle and Lower Slope Morphology
Gullies perpendicular to the margin run from the mud wedge into the middle slope that
contains abundant features of redeposition. Several large-scale channel systems with
fan-shaped terminations occur along the slope. They are of variable lateral extent but
similar in length, typically 15 – 18 km. One of these terminal fans is over 3.3 m wide and
35 m thick (Figure 3). In the north, large blocks and mounds are scattered as far as 20
km into the basin. The mounds are slightly elongated and surrounded by a moat; they
are likely cold-water coral mounds sitting on boulders and large blocks.
Further to the south, the sediment apron with occasional lobes covers 49 km of the
middle-lower slope. The lobes extend up to 18 km into the basin. Regularly spaced
gullies incise the middle slope and act as sediment transport pathway to the basin floor.
45
Figure 3. Multibeam bathymetry of two channelized slope areas with fan-like terminations and an area
with scattered mounds. Inset = Single channel seismic profile across fan deposit at the mouth of a single
channel.
The southernmost area exhibits intensive slumping and mass transport deposition. A
slump scar up to 40 m in height runs for 11 km high up on the margin, indicating a major
margin collapse (Figure 4). The debris traveled as far as 13 km basinward. Blocks are up
to 30 m high and 500 m in diameter. A second mass wasting event indicates repeated
slope failure in this part of GBB.
b
a
Slope Angle
Value
High : 76
Low : 0
Figure 4. (a) Steepness generated from bathymetric map showing high slope angle (~ 76° = yellow line) in
the southernmost area. Slump scars up to 40 m in height are also observed in the lower slope. Debris from
this process are dispersed as far as 13 km into the basin. (b) Three dimensional view of the margin with
onlapping wedge, lower slope scars and debris field.
46
SEDIMENT DISTRIBUTION
Backscatter and sub-bottom profile calibrated with ground-truthed data from northern
part of Great Bahama Bank gives insight into sediment characteristics and their
consistent distribution. The margin and upper slope portion are covered by muddy
sediments of 6-7 km wide. It changes abruptly in the ~20 km wide middle to lower slope,
where it is covered by large blocks of redeposited carbonate, channels, and sediment
apron. The sediment finally transitions into fine-grained basin floor.
POCKMARKS
The northeast dipping slope is dotted by 22 pockmarks with differing morphologies.
They range in diameter from 100 to 2200 m, and depth from 10 to 130 m. There is no
correlation between pockmark diameter and depth due to sediment infill through time
(Figure 5).
Figure 5. (Top) Map of
the pock marks on the
slightly dipping
seafloor. (Below) Plot
of pockmark depth
versus diameter. The
diameter of pock
marks has no
relationship with the
depth. Units in meters.
47
INTERPRETATION AND IMPLICATIONS
The margin parallel distribution of the mud wedge, the apron of coarse to very coarse
redeposited sediment and the muddy basin floor is consistent for over 100 km along the
slope of southwestern GBB. Yet the slope and basin floor morphology varies along strike.
Margin declivity is approximately 30Û but changes abruptly to approximately 76Û in the
southernmost portion of the study area. This sudden increase might be caused by large
scale margin failure.
The moat between the onlapping mud wedge and the margin might be generated by
currents sweeping down and along the margin. Similar moats have been observed in
submersible dives in the southern Exumas.
The sharp transition from mud to a slope apron made up of redeposited sediment has
not been reported from any other locality. The variety of channelized systems, fan
shaped lobes, and debris fields give evidence of the multitude of sedimentary processes
that feed the apron.
The pockmarks in the southwest corner of the study area are located at the outermost
boundary of the Cuban fold and thrust belt. It is likely that migrating fluids and/or
hydrocarbons along the basal detachment are contributing to the high amount of pock
marks that are the surface expression of fluid escape (Sun et al., 2011).
REFERENCES
Malone, M. J., Slowey, N. C., and G. M. Henderson, 2001, Early diagenesis of shallow-water
periplatform carbonate sediments, leeward margin, Great Bahama Bank (Ocean Drilling
Program Leg 166): Geological Society of America Bulletin, v. 113, p. 881-894.
Playton, T. E., Janson, X. and C. Kerans, 2010, Carbonate slopes, in James, N. P., and R. W.
Dalrymple, eds., Facies Models 4: St. Johns, Newfoundland, Canada, Geological Association of
Canada, p. 449-476.
Sun, Q. L., Wu, S. G., et al., 2011, The morphologies and genesis of mega-pockmarks near the
Xisha Uplift, South China Sea: Marine and Petroleum Geology, v. 28, no. 6, p. 1146-1156.
Wilber, R. J., Milliman, J. D. and R. B. Halley, 1990, Accumulation of bank-top sediment on
the western slope of Great Bahama Bank: rapid progradation of a carbonate megabank:
Geology, v. 18, p. 970-974.
48
COMPOSITION OF COLD-WATER CORAL MOUND
“MATTERHORN” AND ITS SURROUNDING SEDIMENTS IN
THE STRAITS OF FLORIDA
Rani Sianipar, Gregor P. Eberli, and Emmanuelle Ducasou1
1)
University of Bordeaux
KEY FINDINGS
ƒ
A 7.03 m core into the 110 m high “Matterhorn” mound retrieved an unlithified
succession of coral floatstone with variable matrix composition.
ƒ
The percentage of corals within the coral floatstone peaks in the fine-grained
matrix unit with a 49.1% coral content.
ƒ
Grain size alternations in the matrix of the floatstone are interpreted to be related
to glacial and interglacial deposition. Corals grow during both times but are more
abundant in the interglacial periods.
ƒ
The mineralogy displays a downcore trend of decreasing aragonite and increasing
low magnesium calcite that is related to early diagenetic changes.
ƒ
Cross-bedded pteropod-foraminifera grainstone sediments surround the
“Matterhorn” giving the sedimentary record of the bi-directional tidal current
regime in the Straits of Florida.
SIGNIFICANCE
The Straits of Florida has
received increasing attention
over the last decades in term of
cold-water coral (CWC) mound
ecosystems (Neumann and
Ball, 1970; Neumann et al.,
1977; Paull et al., 2000;
Grasmueck et al., 2006; Reed
et al., 2006; Roberts et al.,
2006; Correa et al., 2012). New
seafloor mapping technologies
and considerable research
efforts in the Straits of Florida
have documented that a variety
of
factors
influence
the
initiation,
growth
and
distribution of CWC mounds.
Research using an autonomous
underwater vehicle (AUV) and
submersible observations at 3
sites on Great Bahama Bank
(GBB)
slope
(Figure
1)
documented that variability in
Figure 1. Overview of the study area in the Straits of Florida with
white boxes showing location of AUV survey (Modified from
Grasmueck et al., 2007).
sedimentation rates, current regime, and underlying
49
topography control the distribution, development, and morphology of CWC mounds
(Correa et al., 2012). Simultaneously, Rosenberg (2011) measured temperature ranging
from ~4 to 100C, salinity from ~33.6 to 35.3, and seawater density from ~ 27.35 to 27.8
kg/m3 around the CWC in this area.
The sediments within and adjacent to the mounds had only been sampled on the
surface preventing estimates of sediment accumulation rates and recognition of
sedimentary structures. This study fills these gaps by analyzing two cores retrieved
during the CARAMBAR cruise in November 2010 on the slopes of GBB. Both cores were
taken in the area where Correa et al. (2012) identified the largest mound in the AUV data
set, specifically in GBB site 3 west of Bimini (Figure 2).
Figure 2. (B) High-resolution bathymetry map of GBB site 3 western Bimini on the slope of GBB with the
red box outlining the Matterhorn (C) Location of piston cores (CARKS 15 and CARKS 16).
MATTERHON MOUND SITE
The largest mound at GBB site 3, hereafter called the “Matterhorn”, reaches 110 m in
height (Correa et al., 2012). A 7.03 m gravity core (CARKS 15) from the Matterhorn
allows the top portion of the mound to be investigated with regards to internal structure,
composition and growth rate of the mound. A second, 3.26 m long core (CARKS 16) was
taken in the sediments adjacent to the mound. Current data revealed a bidirectional
current regime produced by an internal tide that switches direction from N to S every 6
hours (Grasmueck et al., 2006). The sediment core is expected to record this current
regime in its sedimentary structures.
SEDIMENT CHARACTERISTIC OF THE “MATTERHORN”
Core CARKS 15 from the flank of the “Matterhorn” is basically a coral floatstone with
variable amounts of corals in a matrix of changing composition. The succession can be
divided into five sedimentary units based on the lithology and the different geophysical
and geochemical properties (Figure 3). (A) Coral floatstone in coarse grained matrix, (B)
50
Figure 3. Core photograph, textural description, fossil content and facies and unit boundaries of core
CARKS 15 of the “Matterhorn” on the slope of GBB. Variable amounts of corals are present throughout the
core but the matrix changes in each facies unit.
Coral floatstone with wackestone intercalation, (C) Coral floatstone interbedded with
grainstone, (D) Coral floatstone in fine grained matrix, (E) Lithic rudstone. The matrix of
the coral floatstone is mainly composed of pteropods, planktonic and benthonic
foraminifera, with admixed coral fragments, sponge spiculae, mollusks, and echinoids.
The lithic rudstone unit at the lower portion of the core is mainly composed of micrite
and a few echinoids, planktonic and benthonic foraminifera.
Table 1. Result of coral percentage measurement at certain depths of core CARKS 15 showing a
maximum, minimum and average value of coral content in each facies.
Unit
A
B
C
C
C
D
D
D
D
Depth (cmbsf)
135 – 150
?
342 – 354.6
354.6 – 357
433 – 448
507 – 519.9
519.9 – 522
570 – 585
627 - 642
Facies
Coral floatstone
Coral floatstone
Coral floatstone
Skeletal grainstone
Coral floatstone
Coral floatstone
Skeletal packstone
Coral floatstone
Coral floatstone
Maximum (%)
8.9
14.6
28.9
10
28.9
17.9
6.3
49.1
22.8
Minimum (%)
2
1.7
3.9
6
4
1.8
4.6
8.8
5.4
Average (%)
5.1
6.5
18.6
7
16.7
10.6
4.7
29.4
17.9
51
Coral Quantification
The distribution of the corals in the top meters of the “Matterhorn” is assessed with
tomographic (CT) scan imagery over several intervals of core CARKS 15. Coral density
varies between 1.7% to 49.1%. The highest amount of coral, 49.1%, is in the interval 570 585 cmbsf (cm below seafloor) where corals float within a fine-grained matrix (Table 1).
Geophysical Logging
The Matterhorn core was analyzed with a GEOTEK Multi Sensor Core Logger at a
resolution of 1 cm, measuring natural gamma radiation (cps) and resistivity (mV).
Gamma density and resistivity reflect changes in the lithology and porosity. The Gamma
raw bulk densities measured on core CARKS 15 display relatively parallel trends. The
densities vary between 4,107 to 6,875 cps. The resistivity has values between -1 to 125
mV with an average of 81 mV and displays an inverse correlation with the density
measurement.
Geochemistry
The mineralogy consists mostly of aragonite and low magnesium calcite (LMC), but XRay Diffraction (XRD) documents a decreasing trend of aragonite and increasing trend
of low magnesium calcite (LMC) from top to bottom of the core. In unit A through unit
B, aragonite content varies between 70% to 85%, while LMC varies between 15% to 30%
for sediments of all size fractions. The high percentage of aragonite is in the smallest size
fraction, <63 μm, confirming that the fine mud matrix is likely to be sourced from the
platform top. At a depth of 360 cmbsf in unit C, a significant increase in LMC content is
identified. LMC reach 75% in coral floatstone facies within the size fraction >150 μm,
while the mud matrix still is 84% aragonite. At a depth of 580 cmbsf, however, the finegrained matrix has a relatively low 54-65% aragonite content. In unit D where grains
larger than 150 μm are lithoclasts, aragonite reaches 49% and LMC 41% while the mud
fraction has 32% aragonite and 68% LMC. The decrease of aragonite and the coeval
increase of lithification and LMC document early diagenetic alterations within the
mound.
Down core element intensities of Fe, K and Si, determined with an X-Ray Fluorescence
(XRF) core scanner, are relatively constant downcore except at a depth of 613 to 643
cmbsf, where Fe, K, and Si are enriched, which indicates the presence of clay minerals or
iron-minerals. Visually, this interval is characterized by darker streaks due to presence of
pyrite. Sr intensities are strongly inversely correlated with Fe, K and Si. For example, in
unit D (500 to 600 cmbsf) Sr content increases, while Fe, K, and Si intensities decrease.
The opposite trend can be seen over the lower part of the core at depths of 600 to 660
cmbsf.
SEDIMENT CHARACTERISTICS SURROUNDING “MATTERHORN”
Core CARKS 16 from the side of the “Matterhorn” consists of a coarse-grained skeletal
grainstone to rudstone. The skeletal components are predominantly pteropods,
benthonic and planktonic foraminifera. Portions of the core display large-scale crossbedding. The core can be divided into three lithologic units based on the composition
and depositional texture (Figure 4 & 5).
52
Figure 4. Photograph and lithology of off-mound core CARKS 16, displaying the brownish coarsegrained grainstone-rudstone and the lighter colored packstone to grainstone. The top portion is
massive, while the lower portion of the core displays large scale cross-bedding.
Unit A: Pteropod grain-rudstone (0-113 cmbsf); the light yellowish brown pteropod
grain-rudstone is poorly sorted and displays a coarsening upward trend. The grains
consist of abundant pteropods, planktonic and benthonic foraminifera, and sponge
spiculae. Unit B: Foraminifera grainstone (100-293 cmbsf); the pale brown unit is a well53
sorted silt to very coarse, sand-sized foraminiferal grainstone. The unit is slightly
burrowed but still shows inclined bedding and bi-directional cross-bedding. Unit C:
Foraminifera packstone (294-326 cmbsf); the very pale-brown foraminifera packstone is
moderately sorted, clay to medium sand-sized and composed of mostly planktonic
foraminifera and micrite.
Figure 5. Composition of core CARKS 16. (A) & (B) Pteropod grain-rudstone showing coarse grained
pteropods with finer-grained planktonic and benthonic foraminifera. (C) Foraminifera grainstone
showing abundant planktonic and benthonic foraminifera and some pteropods and echinoids. (D)
Foraminifera packstone with planktonic and benthonic foraminifera in muddy matrix. Pf-planktonic
foraminifera, Bf-benthonic foraminifera, P-pteropods, and E-echinoids.
INTERPRETATION AND IMPLICATIONS
In the Matterhorn core, coral fragments up to 30 mm in length are irregularly
distributed throughout the mound core. The matrix of the coral floatstone, however,
alternates between coarse-grained and fine-grained. The coarse-grained matrix is likely
to reflect times of increased bottom currents such as those associated with lowered sea
level. The fine-grained matrix is mostly composed of aragonite mud that is likely sourced
from the platform top during high sea-level. Both of these matrixes contain abundant
coral fragments but the highest amount is present in an interval characterised by a
muddy matrix. This indicates that cold water corals are growing during both glacial and
54
interglacial times in the Straits of Florida, but living conditions might be slightly better
during interglacials.
The downcore decrease of aragonite content and concomitant increase of LMC is
mostly explained by early diagenesis. The high value of aragonite in the top portion of
core CARKS 15 is from corals and pteropods, while LMC mainly comes from foraminifera
and, to a lesser extent, from mollusks and echinoids. Downcore LMC increases in the
muddy sediment supporting transformation of the metastable aragonite to the more
stable LMC during early diagenesis.
Off-mound core shows layers and lenses of coarse and fine-grained sediments, while
the core from the mound shows alternations of coral-rich and coral-poor layers. The off
mound sediments are mainly composed of coarse pteropod grainstone showing inclined
bedding and bi-directional cross bedding documenting internal tidal current
environment.
REFERENCES
Correa, T. B. S., Grasmueck, M., Eberli, G. P., Reed, J., Verwer, K., and S. Purkis, 2012, Variability
of cold-water coral mounds in high sediment input and tidal current regime, Straits of Florida:
Sedimentology, v. 59, p. 1278-1304.
Foubert, A. and J. Henriet, 2009, Nature and significance of the recent carbonate mound record:
the mound challenger code: Springer-Verlag Berlin Heidelberg.
Grasmueck, M., Eberli, G. P., Viggiano, D. A., Correa, T., Rathwell, G., and J. Luo, 2006,
Autonomous underwater vehicle (AUV) mapping reveals coral mound distribution,
morphology, and oceanography in deep water of the Straits of Florida: Geophysical Research
Letters, v. 33, L23616.
Grasmueck, M., Eberli, G. P. Correa, T. B. S., Viggiano, D. A. , Luo, J., Reed, J. K., Wright, A. E.
and P. A. Pomponi, 2007, AUV-based environmental characterization of deep-water coral
mounds in the Straits of Florida: OTC 18510, Houston, p. 1-11.
Neumann, A. C. and M. M. Ball, 1970, Submersible observations in Straits of Florida - Geology
and Bottom Currents: Geological Society of America Bulletin, v. 81, p. 2861-2873.
Neumann, A. C., Kofoed, J. W., and G. H. Keller, 1977, Lithoherms in Straits of Florida: Geology,
v. 5, p. 4-10.
Rosenberg, A., 2011, Insight from the depth of the Straits of Florida: assessing the utility of
Atlantic deep-water coral geochemical proxy techniques: University of Miami M.S. Thesis,
Open Access Theses, Paper 244, http://scholarlyrepository.miami.edu/oa_theses/244.
Paull, C. K., Neumann, A. C., Ende, B. A. A., Ussler, W., and N. M. Rodriguez, 2000, Lithoherms
on the Florida-Hatteras slope: Marine Geology, v. 166, p. 83-101
Reed, J. K., Weaver, D., and S. A. Pomponi, 2006, Habitat and fauna of deep-water Lophelia
pertusa coral reefs off the Southeastern USA: Blake Plateau, Straits of Florida, and Gulf of
Mexico: Bulletin of Marine Science, v. 78, p. 343-375.
Roberts, J. M., Wheeler, A. J., and A. Freiwald, 2006, Reefs of the deep: The biology and geology
of cold-water coral ecosystems: Science, v. 312, p. 543-547.
55
56
PETROPHYSICAL PERSPECTIVE OF CRETACEOUSTERTIARY RE-DEPOSITED CARBONATES FROM THE
APENNINES AND THE ADRIATIC SEA, ITALY
Irena A. Maura, Gregor P. Eberli, and Daniel Bernoulli1
1)
Geological Institute, University Basel, Switzerland
KEY FINDINGS
Re-deposited carbonates have a good reservoir potential with the following petrophysical
charateristics:
ƒ At the same porosity, re-deposited carbonates have higher permeability and are
faster than the background sediment.
ƒ
In terms of age, Cretaceous re-deposited carbonates are more porous and
permeable than Tertiary ones. Yet, Cretaceous sections have faster velocity
compared to the Tertiary counterparts.
ƒ
The pore structure is dominated by large simple pores producing fast high
permeability rocks. In contrast the pelagic background sediments are dominated
by compliant micropores and are, thus, slower at any given porosity.
CONTROLS ON RE-DEPOSITED CARBONATE RESERVOIR QUALITY
Re-deposited carbonate reservoirs are a challenging prospect for hydrocarbon
exploration. The reservoir quality and potential of these deposits is still questionable.
However, reservoirs in carbonate mass gravity flow deposits exist, indicating that these
deposits can be a good reservoir.
Re-deposited
carbonates
are like their shallow-water
counterparts susceptible to
diagenetic processes that
alter
their
original
mineralogy
and
pore
structure (Eberli et al.,
2003). Diagenetic alterations
create a complex pore
structure that will contribute
to
variations
in
the
petrophysical properties such
as porosity, permeability and
sonic velocity (Weger et al.,
2009). Likewise, the amount
of microporosity also plays Figure 1. (Left) Study sites; the yellow dot indicates location that
an important role in the was visited in 2006, the blue dot indicates the location of Well 1,
various
petrophysical and the red dot indicates locations that were visited in 2012.
properties observed (Baechle (Upper Right) Coarse breccia bed (above back pack) in a series of
thin calcareous turbidites in the Valle di Pennapiedimonte
et al., 2008).
(Maiella). (Lower Right) Monte Conero section.
57
In this study the different petrophysical properties of Cretaceous - Tertiary redeposited carbonates and their background sediment will be assessed in outcrops and
the subsurface. The rock samples were collected from several sites in the Abruzzi and
Apennines in Italy (Figure 1).
DATASETS AND METHODS
Over 210 plugs from the outcrop and 85 core plugs are analyzed. The samples were
collected from three different areas (1) Maiella platform margin, an isolated Mesozoic to
Mid-Tertiary carbonate platform margin located in the southern part of Italy (Figure 1);
(2) Monte Conero, a section located 150 km to the north of the Maiella mountain which
contains mostly turbidites and few breccia beds (Figure 2); and (3) Well 1, an offshore
well in the Adriatic Sea, that penetrated the basinal portions of a buried carbonate
platform.
Each plug was measured for porosity and permeability. In addition, 87 plugs from the
outcrop and 40 core plugs were measured for sonic velocity. Petrographic description
and Digital Image Analysis (DIA) using thin sections were used to examine composition
and pore structure. In addition, Scanning Electron Microscope (SEM) analyses from
some core samples were used to study the composition and micro-pore structure.
Figure 2. (Upper left) Sassi Neri outcrop location (Monte Conero area). (Lower left) Outcrop photo of a
exceptionally thick Cretaceous turbidite bed with flute cast. (Right) Stratigraphic columns of 4 Sassi Neri
outcrop.
58
FACIES OF RE-DEPOSITED CARBONATES
The studied calcareous mass gravity flow deposits, hereafter called “re-deposited
carbonates” are comprised of breccias, turbidites, slumps and calcisiltites. The perennial
background sediment is either periplatform ooze, or, in the distal positions, fine pelagic
sediment. In Well 1 the dominant facies in the Tertiary is megabreccia (28.11 %),
followed by turbidite (12.17 %), whereas in the Cretaceous the facies is dominated by
turbidite (36.59 %) and megabreccia (8.82 %), followed by calcisiltite (7.72 %) (Figure.
3). The carbonate turbidites in Well 1 core are dominated by fine-grained distal
turbidites. A similar
distal
turbiditic
facies is found in
outcrops along the
Marchean
anticlinorium I and
is
the
closest
outcrop to Well 1
(Montanari et al.,
1989).
The
redeposited
carbonates in the
Maiella
outcrop
tend to be coarser
because they are in
a very proximal
position
to
the Figure 3. Facies distribution histogram of Cretaceous – Tertiary re-deposited
carbonate facies from Well 1 (modified from Maura et al., 2011).
platform.
POROSITY, PERMEABILITY AND SONIC VELOCITY
I. Porosity and Permeability of Maiella outcrop versus Well 1 core
Maiella outcrop samples of re-deposited carbonates and the background sediment have
a larger range and higher porosity and permeability compared to Well 1 core data (Figure
4). The higher porosity and permeability of the Maiella samples compared to those from
Well 1 is related to the coarse composition and the proximal location to the source. In
contrast,the Monte Conero samples have a similiar range in porosity (2– 15 %) to Well 1
samples. The re-deposited carbonates tend to be more permeable than the background
sediments (Table 1, Figure 4).
Table 1. Porosity – permeability value of re-deposited carbonates versus background sediments from
the Maiella outcrop and Well 1 core samples.
Dataset
ĭ (%)
Ʈ (md)
Re-deposited carbonates
Maiella
Well 1
Monte
Outcrop
Core
Conero
0.40 - 32
0.4561 - 22
2 – 15
0 - 522
0 - 23
-
Background sediments
Maiella
Well 1
Monte
Outcrop
Core
Conero
1.96 - 27
1.19 - 11
2 – 15
0 - 72
0 - 1.5
-
59
Figure 4. Re-deposited carbonates (left) and background sediments (right) porosity – permeability plot
from Maiella (red dots) and Well 1 (blue dots) samples.
II. Porosity and permeability of Re-deposited Carbonate versus Background Sediment
At a given porosity, re-deposited carbonates tend to be more permeable than the
background sediments. Porosity can reach 32 % with permeability up to 522 md. While
porosity of the background sediment can reach 27 % and permeability 72 md (Table 2).
III. Porosity and permeability of Cretaceous versus Tertiary Re-deposited Carbonate
Both Cretaceous and Tertiary re-deposited carbonates have a good reservoir potential.
In addition, the Cretaceous has a better reservoir quality compared to the Tertiary. The
porosity in Cretaceous re-deposited beds can reach 32 % with 522 md, while in the
Tertiary the highest porosity is 23 % and 17 md permeability (Figure 5). The difference in
reservoir quality is related to the original mineralogy and the resulting diagenetic
potential. During the Cretaceous, calcite was the predominant carbonate mineral, while
in the Tertiary most carbonate production was aragonite. Metastable aragonite is subject
to dissolution and re-precipitation. This process produced more cement and tighter
rocks in the Tertiary.
IV. Sonic velocity
At the same porosity the re-deposited carbonates are faster than the background
sediment but they are more permeable than the background sediment. This discrepancy
is related to the differences in pore structure in the mostly skeletal grainstone to
rudstone in the redeposited beds compared to the dominance of microporosity in the
background sediments.
Table 2. The range of porosity – permeability value of Cretaceous versus Tertiary re-deposited
carbonates and background sediments.
Dataset
ĭ (%)
Ʈ (md)
60
Re-deposited carbonates
Cretaceous
Tertiary
0.40 - 32
1.61 – 23.89
0 - 522
0 – 17.381
Background sediments
Cretaceous
Tertiary
5.19 – 26.95
1.96 – 24.42
0.045 -36.35
0 - 72
Figure 5. (Left) Porosity – permeability plot of Cretaceous and Tertiary re-deposited carbonates and
background sediments. (Right) Thin section photos of A and B in the plot.
Figure 6. Plot of porosity and velocity. Colorbar represents: (A) Perimeter Over Area (POA), (B) Dominant
pore size (DOMsize). We overlay our data with Weger’s (2009) data to compare the trends and the
distributions of the two datasets. Higher DOMsize indicates larger pores, and higher PoA indicates higher
complexity in the pore structure. Large and simple pores are the dominant pore structure of the Maiella
samples.
61
PORE STRUCTURE OF RE-DEPOSITED CARBONATES
Pore structure is one of the important factors that can affect the velocity in carbonates.
Digital Image Analysis (DIA) shows that the majority of pores in re-deposited carbonates
are large pores dominated by simple pores (Figure 6A and B). Such pore structures
produce a fast velocity at a given porosity (Weger et al., 2009)
Another significant factor that affects the velocity behavior in carbonates is the
percentage of microporosity in the rock. Figure 7 shows that at the same porosity,
samples with higher amounts of microporosity are slower. The reason is that the small
pores are composed of soft compliant pores (Baechle, 2008).
Figure 7. Plot of porosity
and velocity. Colorbar
represents the percent of
microporosity. This plot
shows that the samples
with fewer micropores and
lower porosity will have a
faster velocity.
REFERENCES
Baechle, T. G., Colpaert, A., Eberli, G. P., and R. J. Weger, 2008, Effect of microporosity on sonic
velocity in carbonate rocks: The Leading Edge, p. 1012-1016.
Eberli, G. P., Baechle G. T., Anselmetti, F. S., and M. L. Incze, 2003, Factor controlling elastic
properties in carbonate sediments and rocks: The Leading Edge, v.22, p. 654-660.
Maura, I. A., Eberli, G. P., and D. Bernoulli, 2011, Comparison of Cretaceous-Paleocene carbonate
turbidite successions from core and outcrop adjacent to the Maiella platform margin, Italy:
CSL Annual Meeting, p. 107-112.
Montanari, A., Chan, L. S., and W. Alvarez, 1989, Synsedimentary tectonics in the Late
Cretaceous-Early Tertiary pelagic basin of the Northern Apennines, Italy: controls on
carbonate platform and basin development: SEPM Special Publication 44, p. 379-399.
Weger, R. J., Eberli, G. P., Baechle, G. T., Massafero, J. L., and Y. F. Sun, 2009, Quantification of
pore structure and its effect on sonic velocity and permeability in carbonates: AAPG Bulletin,
v. 93, p. 1297-1317.
62
SEDIMENTOLOGY, GEOMETRIES AND LINK TO THE
SUBSURFACE FROM A FIELD-SCALE ANALOG:
THE SIERRA DE LA VACA MUERTA
Michael Zeller, Samuel B. Reid, David L. Giunta1, Ralf J. Weger,
Gregor P. Eberli, and Jose Luis Massaferro1
1)
YPF, Buenos Aires, Argentina
KEY FINDINGS
ƒ
The Quintuco-Vaca Muerta System displays a strong cyclicity that is linked to
eustatic sea-level fluctuations despite some tectonic overprint.
ƒ
Small-scale heterogeneities follow sequence stratigraphic units and boundaries
and can therefore be identified through seismic stratigraphy studies.
ƒ
Synthetic seismic models of both distal and proximal outcrop areas display very
similar geometries to real subsurface seismic data.
ƒ
Local tectonic pulses produce differential developments from the Valanginian
onwards in the outcrop and subsurface areas.
SIGNIFICANCE
The Quintuco - Vaca Muerta System in the Neuquén Basin in Argentina, is considered
to be one of the most promising unconventional plays outside the US. The Vaca Muerta
shale represents the source rock for most petroleum systems in the basin and its
thickness (up to > 1km) and high TOC values make it a prime target for unconventional
hydrocarbon exploration.
In order to identify best exploration and production techniques, the stratigraphic
architecture as well as the bed-scale heterogeneities have to be understood. They are best
studied in field-scale outcrop analogs. The Sierra de la Vaca Muerta (SdlVM) represents
a key study area with excellent exposures displaying the architecture of the mixed
carbonate siliciclastic system, while being easy to access and tectonically relatively
undisturbed. Using the detailed facies descriptions from the outcrop and a geometrical
framework derived from high-resolution satellite imagery, a new large-scale correlation
provides insights into the sequence architecture and places the observed small-scale
heterogeneities within this framework.
This information is most useful, if directly linked to the subsurface exploration fields.
Synthetic seismic modeling in combination with 2D lines acquired across the study area
provide the connection to subsurface studies and can help to understand the subsurface
architectural elements and stratigraphic composition, which play an important role in
unconventional evaluation of, for example, carbonate content and brittleness.
METHODS
This study is an integrated approach to document the outcrop facies distribution and
geometries, illustrate their seismic expressions, and correlate between the outcrop area
and the subsurface producing fields. Therefore the workflow is subdivided into 3 main
63
portions: 1) Stratigraphic Architecture, 2) Synthetic Seismic Model, and 3) Seismic
Outcrop – Subsurface Correlation.
Stratigraphic Architecture
Figure 1 illustrates the applied workflow, which was used to develop the field-scale
facies correlation. During fieldwork, detailed sedimentological descriptions are collected
in both vertical and lateral sections in order to document spatial and temporal variations
of the depositional system. All points of observation are recorded with GPS coordinates
and facies types are defined and grouped into facies associations.
In a second step, newly acquired high-resolution satellite imagery (Worldview II, 0.5 m
resolution) is draped on top of digital elevation model (DEM, ASTER, 15m resolution) in
order to build a photorealistic digital outcrop model (DOM). The high quality images
together with the correct elevations allow tracking of 49 beds within a 1km thick
stratigraphic section. This information is used to construct a geometry model and
calculate true stratigraphic thickness based on dip and distance measurements.
Finally, observed color transitions on the satellite imagery together with information of
the topographic dip angles from the DEM, in combination with the logged beds and
sections, facilitate interpretation of facies and stratigraphy in inaccessible areas.
Petrophysics and Synthetic Seismic Modeling
The synthetic seismic modeling follows the same workflow that was successfully
applied at the Picún Leufú Anticline (Zeller et al., 2011a). In the first step, the facies
model (Figure 2) is used as the geometrical input and is placed at 2km depth. 532 facies
bodies are defined and 333 traces along the 10km section ensure sufficient lateral
resolution.
Petrophysical properties of 70 samples, covering all facies types, are measured. They
include dry vp, dry vs (both from 5 – 60 MPa), bulk and grain densities, porosity and
carbonate content. The velocity values (for 50MPa ~ 2km burial depth) are corrected
from dry to realistic wet state using the Gassmann fluid substitution equations. The
average values determined for each facies (Table 1) are then distributed according to the
facies model. The resulting acoustic impedance model (based on the velocity and density
input) is then convolved with a 5-10-50-60 Butterworth wavelet and models for 0,10, 20
and 30 degree incident angles are stacked to create a synthetic seismic secti0n that can
be compared with real subsurface datasets.
Table 1. Average petrophysical properties for different facies associations after Gassmann
Fluid Substitution. These values are used for property distribution within the geometry model
and basis for the synthetic seismic modeling. vp and vs values for 50 MPa effective pressure.
facies
vp(m/s)
vs(m/s)
ʌ(g/cm3)
Ɍ(fract)
Carb(fract)
facies
vp(m/s)
vs(m/s)
ʌ(g/cm3)
Ɍ(fract)
Carb(fract)
64
shale
3547
2109
2.25
0.07
0.20
WS
5652
3090
2.63
0.02
0.72
calc.shale
4370
2518
2.45
0.04
0.51
PS
5365
2972
2.56
0.04
0.66
calc.silt
4676
2687
2.53
0.07
0.60
oysterFS
5650
3077
2.58
0.05
0.74
calc.sand
4849
2830
2.51
0.07
0.36
GS
5375
2934
2.56
0.06
0.81
puresand
4215
2623
2.34
0.11
0.02
deltaicsand
4174
2466
2.38
0.10
0.17
Seismic Outcrop Subsurface Correlation
During exploration a fairly dense grid of 2D seismic lines was shot also throughout the
western portions of the Neuquén Basin, covering large parts of the study area. 2D lines
from the western area are tied to the 3D volumes in the eastern portion of the basin and
allow direct connection between outcrop and subsurface areas. A key reflection within
these regional lines reaches the surface in the study area, which allows correlation of this
reflector to the existing digital outcrop model.
Figure 1. Outcrop correlation workflow. Geometry and facies distribution are based on the combination of
field observations and digital outcrop model interpretations.
65
66
Figure 2. Facies architecture and distribution at the SdlVM, based on logged sections, mapped beds and analysis of geometries from high-resolution
satellite imagery and the digital elevation model. Approximately 10km long and along the maximum dip (South to North).
67
Figure 3. Heterogeneities in the SdlVM (Locations marked top right). A) mapped carbonate bed (Bed D in Figure. 1); B) sigmoidal unidirectional prograding beds in contrast to typical tidal structures; C) downlapping oyster floatstone; D) dm-scale mixed turbidite; E)
typical m-scale shale cycles (differences in carbonate content); F) oyster buildups reaching several meters; G) Los Catutos Member, cyclic
marls –limestone.
RESULTS AND INTERPRETATIONS
Large-Scale Architecture
Lithologically the studied interval represents a truly mixed succession. Both pure
carbonate and pure siliciclastic portions are very scarce. Overall the package follows a
prograding/shallowing upward trend with lower basinal deposits overlain by shelfal
sediments that finally shifts rapidly into deltaic deposits (Figure 2). This long-term trend
is repeatedly interrupted by marine transgressions on the shelf. The succession is
subdivided into 8 depositional sequences (Figure 2, I-VIII). The sequences are
commonly asymmetric with thick regressive portions on the shelf, while this pattern is
slightly alleviated in the basinal portions with the thickening of the transgressive
portions.
Depositional geometries are mostly gentle with dip angles around 0.5-1°, with the
exception of sequence V, where a pure carbonate system produces a strong break along
the shelf and a depositional dip of around 3° along its lower portion.
The largest shift of the depositional environment occurs within sequence VIII with the
onset of an unidirectional prograding deltaic facies (Figure 3B) on top of the otherwise
tidally dominated shelfal succession. This strong change could have been associated with
a change in subsidence rate and tectonic movement in the hinterland area, increasing the
siliciclastic input to the study area.
Small-Scale Heterogeneities
Within this large-scale facies architecture the outcrops at the SdlVM also allow more
detailed insights into the smaller scale heterogeneities of the mixed system. These
include cyclic variations, lateral facies transitions and event beds.
The Los Catutos Member (Figure 3G) is a meter scale alternation of skeletal
wackestones with marls (Figure 3E) . Shale successions follow the same pattern and
display on a meter scale variations of more and less carbonate content. These cycles are
laterally associated with more carbonate-rich systems on the shelf. Similar limestone –
marl successions in the northern part of the Neuquén Basin have been associated with
orbital cyclicity and interpreted as regressive carbonates covered by transgressive marl
layers (Kietzmann et al., 2011).
Lateral facies transitions occur in both clastic and carbonate dominated environments.
As illustrated in Figure 3A (see also Zeller et al., 2011b), the high energy carbonate
system, consisting of oolitic grainstone and skeletal packstone shifts laterally into a
mixed shelfal system and finally into slope facies (similar to Los Catutos) and basinal
calcareous shale. Likewise, siliciciclastic sandstones develop downdip to sand-siltstone
mixtures into pure shale deposits along several kilometers.
Oyster float- and framestone are the most common carbonate dominated facies in the
system and can occur as extensive dm-m beds (Figure 3C) or as up to 4m buildups. One
particular buildup is mapped out and shows rapid thickening from 0.1 to more than 3m
(Figure 3F).
In addition to the cyclic and lateral variations, some sedimentary structures provide
evidence for sedimentation events during the time of deposition. Soft sediment
deformation is a common feature within the regressive portions of siliciclasticdominated sequences (Sequences VI-VIII) and can be explained by the rapid
sedimentation of silt and sand on top of the shale. In contrast mixed carbonate clastic
turbidites (Figure 3D) can be found just above carbonate dominated sequence tops
68
(sequences I and IV) and might have been triggered by the steeper depositional profile
built by the precursor carbonate systems.
Figure 4. Petrophysics and synthetic seismic model. A) Vp vs. porosity, color-coded with carbonate content,
B) vp/vs vs. porosity, color-coded with carbonate content. Both plots illustrate that porosity combined with
the carbonate content exert the main control on the sonic velocities. C) Synthetic seismic model based on
the acoustic impedance model from facies distribution, the petrophysical measurements at 50 MPa.
Petrophysics and Seismic Expressions
Porosity and carbonate content have the strongest impact on sonic velocity in the
mixed system of the Vaca Muerta (Figure 4A). This finding corroborates the results from
the more proximal succession at the Picún Leufú Anticline (PLA) (Zeller et al., 2011a),
Overall carbonate-rich facies show higher velocities and higher vp/vs ratios and are
commonly less porous than clean siliciclastics due to cementation (Figure 4A-B).
Carbonate content has a particularly strong effect on shale properties with calcareous
69
shale having considerably higher sonic velocities than their pure shale equivalents (Table
1).
In the resulting synthetic seismic model the general character of the prograding –
aggrading succession is well preserved (Figure 4C). In the lower portion the beds
downlap on the Los Catutos Member. The strongest geometrical anomaly is represented
by the carbonate succession of sequence IV. The following beds onlap onto this key
reflection and are overlain by a more aggradational succession, which exposes quite
commonly lateral variations in amplitudes and are very well-defined as downlapping
beds in the lower portion.
Identification of depositional sequences is not as straight forward as from the facies
correlation. Some of the thin sequences would most likely be missed in seismic
interpretation of real subsurface seismic data with similar resolution.
IMPLICATIONS
Sequence Stratigraphy
In outcrop the succession can be subdivided into 8 depositional sequences. Based on
biostratigraphic studies (Leanza et al., 2011) the studied interval extends from the Early
Tithonian to the Middle – Late Berriasian, a time span of approximately 10 Myr. In
subsurface seismic data 7 sequences are identified that correspond to the 7 deepening
and shallowing cycles represented in the eustatic sea-level curve proposed by Mitchum
and Uliana (1985). The similar number of sequences indicates a similar stratigraphic
framework for both outcrop and subsurface, which are controlled by eustatic sea-level
changes. This is important, since the value of the outcrop as an analog is greatly
enhanced if both areas were controlled by similar depositional processes.
Heterogeneities
Small-scale variations of depositional facies follow lateral and sequence stratigraphic
trends. Even event beds like soft sediment deformation structures and turbidites occur in
predictable intervals within the sequence stratigraphic framework. This could have a
considerable impact on identification of potential sweet spots for unconventional
exploration. Intervals with higher carbonate content represent best spots for hydraulic
fracking, while intervals with high turbidite frequency would offer relatively elevated
porosity and permeability streaks within an otherwise tight surrounding.
Regional Seismic Architecture (Outcrop – Subsurface)
With now two synthetic seismic models in place, both in proximal (PLA) and distal
(SdlVM) positions, comparisons of the regional architecture of outcrop and subsurface
hold very interesting insights.
Figure 5 is a combination of a regional subsurface line from the producing fields
towards the NW and the two synthetic seismic datasets from this study. Placed with the
same scales, they show very similar characteristics. Proximal Portions (PLA and E
subsurface) are relatively thin, have reflectors with gentle dips up to 1° and no breaks
along their profile. In contrast distal portions (SdlVM and W subsurface) are very thick,
both show steeper dip angles in the lower portion and contain reflections with clear
breaks.
Again, these similar developments enhance the value of the outcrops as analogs and
point to similar depositional processes and environments in the different areas over
most of the studied interval.
70
Figure 5. Comparison subsurface with synthetic seismic data from outcrop areas. Top: Regional seismic
line (modified from Leanza et al., 2011); Base: synthetic seismic sections from this study, in same scale as
the subsurface dataset.
Differential Development Outcrop Subsurface (Quintuco vs Quintuco sensu strictu)
Based on the direct outcrop subsurface correlation, marginal marine (Picún Leufú
Anticline) and deltaic deposits (SdlVM) are time equivalent to widespread mixed
carbonate clastic cycles in the subsurface area (Figure 6). These observations suggest
that the western (outcrop) portions were subject to a major siliciclastic input due to a
tectonic uplift in the south, while the eastern areas remained tectonically rather calm in
their shallow marine setting. This finding could help to explain the major discrepancies
between the (outcrop) Quintuco s.s. and the subsurface time-equivalent Quintuco
Formation, which was for a long time a major topic for discussions on the applicability of
the outcrops as analogs for the subsurface.
71
.
Figure 6. Comparison of outcrop and subsurface lithologies. A) Regional outcrop correlation (modified
from Leanza et al., 2011), B) Regional subsurface model based on subsurface data (Mitchum and Uliana,
1985). Dotted red line marks the time equivalent surface in outcrop (A) and subsurface (B).
REFERENCES
Kietzmann, D. A., Martin-Chivelet, J., Palma, R. M., Lopez-Gomez, J., Lescano, M., and A.
Concheyro, 2011, Evidence of precessional and eccentricity orbital cycles in a Tithonian source
rock, The mid-outer carbonate ramp of the Vaca Muerta, northern Neuquen Basin, Argentina:
American Association of Petroleum Geology Bulletin, v. 95, no. 9, p. 1459-1474.
Leanza, H., Sattler, F., Martinez, R. S., and O. Carbone, 2011, La formacion Vaca Muerta y
equivalentes (Jurassico Tardio – Cretacico Temprano) en la Cuenca Neuquina. In: Leanza, H.,
Arregui, C., Carbone, O., Danieli, J. C., and Valles, J. M., Geologia y recursos naturales de la
provincia del neuquen: Relatorio del XVII Congreso Geologico Argentino, p. 113-130.
Mitchum, R. M., and M. A. Uliana, 1985, Seismic stratigraphy of carbonate depositional
sequences, Upper Jurassic-Lower Cretaceous, Neuquen Basin, Argentina. In: Bero, B. R., and
D. G. Wooverton, 1985, Seismic stratigraphy: an integrated approach to hydrocarbon
exploration: American Association of Petroleum Geology Memoir 39, p. 255-274.
Zeller, M., Weger, R. J., Eberli, G. P., Giunta, D. L., and J. L. Massaferro, 2011a, Seismic
expressions of a field-scale Quintuco Vaca Muerta outcrop analog – implications for seismic
interpretation of mixed carbonate-siliciclastic systems: CSL Abstracts 2011, p. 61-68.
Zeller, M., Reid, S. B., and Eberli, G. P., 2011b, The distal mix – shelf to basin transitions of the
mixed Quintuco-Vaca Muerta System: CSL Abstracts 2011, p. 69-73.
72
DECOUPLED INORGANIC AND ORGANIC CARBON ISOTOPE
RECORDS: A GLOBAL SIGNAL UNRELATED TO GLOBAL
CARBON CYCLING?
Amanda M. Oehlert and Peter K. Swart
KEY FINDINGS
ƒ
The carbon isotope composition (G13C) of pelagic carbonates deposited on the
Walvis Ridge in the south Atlantic do not show a positive covariation through
time, contrary to theoretical expectations.
ƒ
The G13C of the organic and inorganic fractions from carbonate slope sediments
show different relationships depending on platform margin architecture.
-
Inorganic and organic G13C records in a transect of Ocean Drilling Project
(ODP) cores off the margin of the Great Barrier Reef shows no significant
relationship through time.
-
ODP cores from the Great Australian Bight show a variable relationship
between inorganic and organic G13C records through time.
-
Compared to published results from the Great Bahama Bank, each of
these three slopes exhibits different relationships, suggesting that local
depositional processes and platform architecture significantly influence
the G13C records.
SIGNIFICANCE
Synchronous excursions in inorganic and/or organic G13C records from globally
distributed basins have been used to create chemostratigraphic correlations in a variety
of carbonate deposits. Carbon isotope chemostratigraphy of either inorganic or organic
G13C records is especially useful in deposits where other age dating techniques lack
resolution. Of major concern, however, is the impact of open-system diagenesis, where
the isotopic composition of the bulk inorganic carbonate is subjected to significant
isotopic alterations. One method used to prove the primary nature of the inorganic G13C
record is to conduct a paired carbon isotope analysis. A positive covariation between
inorganic and organic G13C records has been used as a tool to prove that the bulk
inorganic G13C isotopic record represents changes in the global carbon cycle. This
analysis has been applied to both shallow and deep marine settings in studies of global
carbon cycling in time periods spanning the Proterozoic to the present; however, the
ability of the shallow marine and deep marine settings to record the same trends has not
been evaluated. An evaluation of this assumption needs to be conducted so that
interpretations and stratigraphic correlations of G13C excursions accurately reflect the
processes that generate the bulk G13C signal.
In order to test this assumption, this study has quantified the relationship between
inorganic and organic G13C records in both a pelagic setting, and in two platform-to-basin
transects of ODP cores. The pelagic relationship, analyzed in pelagic carbonates from
73
Walvis Ridge, was then compared to relationships produced from new records at the
Great Barrier Reef and the Great Australian Bight, and published records from the Great
Bahama Bank. In each of these cases, a variable relationship between inorganic and
organic G13C values was observed, suggesting that local controls significantly influence
the bulk isotopic composition of inorganic and organic carbon in these environments. As
a result, a knowlege of the processes that contribute to the bulk isotopic composition of
organic and inorganic G13C records will improve interpretations of the significance of
excursions in the G13C values of marine carbonates and sedimentary organic matter. A
better understanding of the significance of the excursions may aid in chemostratigraphic
correlations.
Figure 1.
(A) Location of DSDP
Site 525 (From Bergren et
al., 2003).
(B) Relationship between
inorganic and organic
carbon isotope records
from Site 525.
RESULTS
Pelagic Setting: Walvis Ridge in the South Atlantic, DSDP Site 525
DSDP Site 525 is located on the Walvis Ridge in the South Atlantic (Figure 1a) and has
been described as a sequence of well-preserved pelagic carbonates. These sediments
were studied in 1984 by Shackleton and Hall who related a trend in the carbon isotope
record to a progressive increase in the transfer of organic carbon to the inorganic carbon
reservoir in what has become a seminal paper in global carbon cycling research. The data
produced in the Shackleton and Hall (1984) paper is the foundation for quantitative
carbon cycling models through time (Kump and Arthur, 1999; Shackleton, 1985), which
suggests that the relationship between inorganic and organic G13C values at the DSDP
Site should exhibit a positive covariation through time as predicted by theoretical
fractionation calculations. However, the results of the paired isotope analysis conducted
74
during this study at Site 525 revealed no significant correlation between inorganic and
organic G13C records (Figure 1b).
Figure 2. Location of cores analyzed from ODP Leg 182. From Feary et al., 2003.
Periplatform Setting: Great Australian Bight, ODP Leg 182
The Great Australian Bight is a sub-tropical, cool water carbonate factory located on the
southern coast of Australia. ODP Sites 1128, 1134, and 1132 were selected for paired
carbon isotope analysis (Figure 2). These sites are aligned in a roughly proximal to distal
transect, spanning the platform margin to basin transition south of the Outer Eucla
Shelf. Of the three sites selected for analysis, ODP Site 1132 is the most proximal drill
site, while ODP Site 1128 is the most distal. The results of the paired inorganic and
organic G13C record analysis showed that the relationship between inorganic and organic
G13C values is variable in this transect.
Periplatform Setting: Great Barrier Reef, ODP Leg 133
ODP Leg 133 was drilled off the eastern coast of Australia, near Cairns (Figure 3a). The
Great Barrier Reef margin is a mixed siliciclastic-carbonate, reef-rimmed margin. Three
sites (820, 823 and 811) located in a basin-to-platform transect were selected for this
analysis. No significant relationship between inorganic and organic G13C records was
observed at any location (Figure 3b).
INTERPRETATION AND IMPLICATIONS
The results of this study highlight the importance of local depositional processes and
global sea-level variability. While some of these variables, like eustatic sea-level changes,
exert a global influence on carbonate slopes around the world, processes like
depositional patterns and margin architecture seem to play a large role in determining
the isotopic signature in the bulk inorganic and organic G13C records. However, many
75
Figure 3. Relationship between inorganic and organic carbon isotope records at the Great Barrier Reef
transect.
published studies use a positive covariation between inorganic and organic G13C records
to prove the global nature of the inorganic record (Swanson-Hysell et al., 2010; Johnston
et al., 2012; Fischer et al., 2009; Grotzinger et al., 2011). Furthermore, these studies have
also used a positive correlation to reconstruct ancient atmospheric concentrations of
carbon dioxide (Hayes et al., 1999; Rothman, 2002; Jasper and Hayes, 1994; Fike et al.,
2006). In many of these cases, the local processes that seem to play an important role in
determining the composition of the inorganic and organic G13C are not considered.
Theoretical calculations of the fractionation of dissolved inorganic carbon by
photosynthesis suggest that the inorganic and organic carbon fractions should be offset
by a consistent and predictable fractionation factor. As a result, excursions in the
isotopic composition of organic and inorganic carbon should be simultaneous. This
would result in a positive covariation between inorganic and organic G13C records
through time. However, the results of this study suggest that inorganic and organic G13C
records derived from pelagic carbonates may not always co-vary through time as
expected. The inorganic and organic carbon isotope records produced from pelagic
carbonates at Site 525 show no relationship (r2=0.08). Interestingly, the range in
organic G13C values (-28 to -21‰) at Site 525 may provide some evidence of terrestrial
organic matter contributions to the deposit. Pelagic organic material typically falls
around -21‰ (Laws et al., 1995), while some types of terrestrial organic material are
more isotopically depleted. The depleted organic G13C values measured in the sediments
from Site 525 may imply that even pelagic sequences may be considered a mixed system,
depending upon the regional oceanographic circulation.
76
Figure 4. Relationship between inorganic and organic carbon isotope records along the basin-to-profile
transect of the leeward margin of Great Bahama Bank (From Oehlert et al., 2012).
Shallow marine and periplatform carbonates have been substituted for pelagic G13C
records in geologic strata older than 200 Ma. In the majority of cases, pelagic carbonates
older than 200 Ma have been subducted, making it necessary to find another source of
inorganic and organic carbon to analyze in order to reconstruct the ancient global carbon
cycle. This substitution has not been evaluated, and the ability of periplatform
carbonates to record the same type of information about global carbon cycling has not
yet been tested. Recently, it has been shown that the relationship between inorganic and
organic G13C records can be more complicated than previously thought. Within a 30 km
basin-to-platform transect, the relationship between inorganic and organic G13C varied
between a strong positive correlation (r2=0.64) in the basin to no relationship at the top
of the slope (Oehlert et al., 2012, Figure 4). Therefore, along with the analysis of a pelagic
carbonate setting at DSDP Site 525, the analyses of the relationship between inorganic
and organic G13C records in slope carbonates conducted during this study provide a
quantitative appraisal of whether or not the slope carbonates G13C values record the same
information as the pelagic G13C values.
The results of the analyses on the slope carbonates from the Great Australian Bight and
that Great Barrier Reef agree with the findings of Oehlert et al. (2012) in that the
relationship between inorganic and organic G13C records at a cool-water carbonate
factory and a mixed clastic-carbonate margin were observed to be variable. These results
suggest that local factors override the signature of the global carbon cycle in carbonate
slope environments. Local factors such as depositional processes, margin architecture,
mineralogy of the inorganic carbon, and the variability in the organisms that are
contributing organic carbon to the slopes are important factors to consider when
interpreting the significance of excursions in inorganic and organic G13C records sourced
from shallow marine carbonates.
Finally, the inconsistency in the relationship between inorganic and organic G13C
records in both environments suggests that using paired isotope analyses to prove the
77
original nature of a bulk inorganic G13C record should be reconsidered. From the results
of these analyses, sedimentological and sequence stratigraphic characteristics of the
deposit are key variables that should be incorporated into any interpretation of the
significance of excursions in G13C records.
REFERENCES
Fike, D. A., Grotzinger, J. P., Pratt, L. M., and R. E. Summons, 2006, Oxidation of the Ediacarn
ocean: Nature, v. 444, p. 744-747.
Fischer, W. W., Schroeder, S., Lacassie, J. P., Beukes, N. J., Goldberg, T., Strauss, H., Horstmann,
U. E., Schrag, D. P., and A. H. Knoll, 2009, Isotopic constraints on the Late Archean carbon
cycle from the Transvaal Supergroup along the western margin of the Kaapvaal Craton, South
Africa: Precambrian Research, v. 169, n. 1-4, p. 15-27.
Grotzinger, J. P., Fike, D. A., Fischer, W. W., 2011, Enigmatic origin of the largest-known carbon
isotope excursion in Earth’s history: Nature Geoscience, v. 4, p. 285-292.
Hayes, J. M., Strauss, H., Kaufman, A. J., 1999, The abundance of 13C in marine organic matter
and isotopic fractionation in the global biogeochemical cycle of carbon during the past 800
Ma: Chemical Geology, v. 161, p. 103-125.
Jasper, J. P., Hayes, J. M., Mix, A. C., and F. G. Prahl, 1994, Photosynthetic fractionation of 13C
and concentrations of dissolved CO2 in the central equatorial Pacific during the last 255,000
years, Paleoceanography, v. 9, n. 6, p.781-798.
Johnston, D. T., Macdonald, F. A., Gill, B. C., Hoffman, P. F., and D. P., Schrag, 2012, Uncovering
the Neoproterozoic carbon cycle: Letters to Nature, v. 483, p. 320-323.
Kump, L. R. and and M. A. Arthur, 1999, Interpreting carbon-isotope excursions: carbonates and
organic matter: Chemical Geology, v. 161, p. 181-198.
Laws, E. A., Popp, B. N., Bidigare, R. R., Kennicutt, M. C., and S. A. Macko, 1995, Dependence of
phytoplankton carbon isotopic compositions on growth-rate and [CO2] (Aq)-theoretical
considerations and experimental results: Geochimica, Cosmochimica Acta, v. 59, p. 1131-1138.
Oehlert, A. M., Lamb-Wozniak, K. A., Devlin, Q. B., Mackenzie, G. J., Reijmer, J. J. G., and P. K.
Swart, 2012, The stable carbon isotopic composition of organic material in platform derived
sediments: implications for reconstructing the global carbon cycle: Sedimentology, v. 59, p.
319-355.
Rothman, D. H., 2002, Atmospheric carbon dioxide levels for the last 500 million years: PNAS, v.
99, n. 7, p. 4167-4171.
Shackleton, N. J., 1985, Atmospheric carbon dioxide, orbital forcing, and climate. In: The Carbon
Cycle and atmospheric Co2: natural variations Archaen to Present: Geophysical Monograph
(Eds. E. T. Sundquist and W. S. Broecker), v. 32, p. 412-417.
Shackleton, N. J. and M. Hall, 1984, Carbon Isotope Data from Leg 74, In: Initial Reports Deep
Sea Drilling Project (Eds. J.T. Moore and P. Rabinowitz), v. 74, p. 613-619.
Swanson-Hysell, N. L., Rose, C. V., Calmet, C. C., Halverson, G. P., Hurtgen, M. and A. C. Maloof,
2010, Cryogenian glaciation and the onset of carbon-isotope decoupling: Science, v. 328, p.
608-611.
78
APPLICATION OF CAVITY RINGDOWN SPECTROSCOPY TO
STABLE ISOTOPIC MONITORING OF CO2 SEQUESTRATION
DURING ENHANCED OIL RECOVERY
Benjamin T. Galfond, Daniel D. Riemer, and Peter K. Swart
KEY FINDINGS
ƒ
Routine monitoring of the concentration and G13C of gases can be accomplished
under field conditions using a Cavity Ring Down Spectrometer (CRDS).
ƒ
The CRDS is superior to traditional Isotope Ratio Mass Spectrometry in that
there is rapid access to data enabling real time decisions to be made based on the
changes in concentration and G13C values.
ƒ
Even subtle variations in the concentration and composition of CO2 due to
biogenic activity can be reliably measured.
BACKGROUND
Atmospheric concentrations of carbon dioxide (CO2) have rapidly increased over the
past two centuries as a result of the burning of fossil fuels. One approach to deal with the
increase is to store or sequester the CO2 underground. With viable injection sites all over
the globe, millions of tons of gas may be stored for millennia in underground reservoirs,
and when coupled with practices such as enhanced oil recovery this can even be an
economical process.
One question which must be addressed is whether the CO2 remains within the storage
site. Geochemical observations of CO2 concentrations provide an insight into possible
leaks, although with a large number of natural sources, flux measurements may not
always represent reservoir leakage. By examining the stable isotopic signature of the
emitted CO2, we can better attribute increased soil gas emissions to their sources, be they
natural or a result of the sequestration process. When coupled with the other geological
data, we can better attribute the CO2 signature we see near the surface with subterranean
seismic activity and gas migration.
Traditional stable isotope ratio mass spectrometry (IRMS) is poorly suited to high
temporal resolution in-situ studies, because of size, power, and cryogen requirements.
In contrast, a new optical technique, Cavity Ringdown Spectroscopy (CRDS), has the
potential to play an important role in such measurements (Figure 1) because of the
instrument’s small size and low power requirements.
INSTRUMENT DEPLOYMENT
A site was chosen in Texas, where CO2 derived from a natural source (Jackson Dome)
was being injected to support enhanced recovery. In order to monitor potential leakage
a range of analytical methods were deployed at the site including synthetic aperture
radar (SAR), Global Positioning System (GPS), and Seismic analysis. Deployment of the
CRDS to the site occurred in February 2012. The instrumentation setup consists of a
79
centrally located shed that houses the bulk of the equipment and 13 sampling positions
spread radially approximately 150 m from the center covering an approximate 70,000
m2 footprint, allowing for topographical constraints (Figures 2 and 3). A PVC pipe is
sunk two feet into the soil (Figure 4) and terminates above the surface in a polyurethane
tube that runs back to a vacuum manifold located inside the shed. All tube lengths are
kept identical in order to ensure an equal distribution of vacuum between the lines.
Figure 1. The CRDS in the lab at RSMAS prior to
deployment at the Texas site.
Figure 2. Sampling manifold which allows the
sampling of 13 locations plus three standards.
Figure 3. Sampling locations. Analyzer is located
near position 1.
Figure 4. Sampling point
Constant suction by a KNF rotary pump ensures that the sample taken from each line is
a current representation of soil gas conditions rather than an amalgam of stagnated
gasses remaining since the last time the line was sampled. Prior to entering the vacuum
manifold, each line is tapped into by a VICI Valco 16 position valve (Figure 2). The
electrically actuated valve allows for the automated selection of any sampling position.
The valve feeds sample gas directly to the Picarro CRDS on site and is programmed and
actuated by the system as well. This system is housed inside a shed on site. A steel space
frame covered in pressure treated plywood provides a stable, dry foundation for the
structure. Power is wired directly into the shed from nearby overhead lines, with one
circuit devoted to a power conditioning battery backup unit and a second circuit for
cooling purposes.
As a result of the remote nature of the selected site, anthropogenic contributions from
industry and transportation should be at a minimum. The predominant signal aside
from any potential leakage would thus be from the local plant and soil microbial
community.
In order to better assess such contributions, we have begun a
80
comprehensive assessment of the area soils and vegetation for carbon and nitrogen
content and isotopic composition. By combining the range of varied data being collected
at the site with geochemical modeling we are able to constrain the measurements
obtained by the CRDS instrument.
Figure 5. The inversely correlated diurnal patterns of CO2 concentration and isotopic composition over
the course of one week are easily identified.
INTERPRETATION AND IMPLICATIONS
With the high temporal resolution provided by the CRDS instrument, we could
observe diurnal trends in CO2 uptake and emission by plant and microbial sources, as
well as seasonal shifts in biogenic activity (Figure 5). When local weather data is crossreferenced with CO2 values, the influence of precipitation can be observed as well. The
isotopic values of the CO2 contributed by these sources was monitored and recorded over
time, helping to develop a baseline of expected emissions against which an emission
from Enhanced Oil Recovery (EOR) related activities would be more easily identified.
With ambient CO2 į13C value at ~ -8‰ and the biological and anthropogenic
contributions far more depleted (-20 to -30‰), the G13C signature of the injectant at
about -3.6‰ to -2.6‰ (Zhou et al., 2003) is easily identifiable. An above average
methane concentration can also be found in the injectant and can be measured by our
field instrument. An injectant emission event would thus be characterized by an increase
in CO2 and CH4 concentration, enrichment of the CO2 isotopic composition, and a
Keeling plot of delta value vs. 1/CO2 concentration yielding an intercept consistent with
the G13C of the injectant. An expected emission event during a routine site maintenance
activity provided an excellent opportunity to verify the integrity of our detection system.
Event detection and analysis was conducted prior to being informed of the gas release.
As expected, we saw an increase in concentration of the species of interest and a
positively trending isotopic composition. When further analyzed, we found the
contributing gas to have an isotopic composition of -3.9‰. As a mix of both the
81
injectant gas and the more negative biogenic sources that characterize a normal
background emission, this is an excellent match to the injectant being released (Figure
6).
Figure 6. Increased CO2 concentration mirrors an isotopic enrichment. This excess gas matches the
isotopic fingerprint of the injectant at -3.9‰
CONCLUSIONS
CRDS is an emerging technology that is already in a position to supplant traditional
IRMS for high resolution field monitoring analysis. Though there are many factors to
consider when deploying instrumentation remotely, this system can produce reliable
continuous measurements with minimal required on-site interaction. The data from
such a system is able to observe and easily identify the release of injectant CO2, as well as
the more subtle diurnal and seasonal variations in ambient CO2. As a key component of
a multifaceted geophysical and geochemical monitoring regimen, CRDS is a proven asset
in the monitoring of carbon sequestration operations.
REFERENCES
Zhou, Z., Ballentine, C. J., Schoell, M., and S. H. Stevens, 2003, Noble gas tracing of subsurface
CO2 origin and the role of groundwater as a CO2 sink: American Geophysical Union, Fall
Meeting, v. V51H-0379.
82
SUB-MICRON DIGITAL IMAGE ANALYSIS (BIBSEM-DIA),
PORE GEOMETRIES AND ELECTRICAL RESISTIVITY IN
CARBONATE ROCKS
Jan H. Norbisrath, Gregor P. Eberli, Ralf J. Weger, Klaas Verwer,
Janos Urai, Guillaume Desbois, and Ben Laurich
KEY FINDINGS
ƒ
DIA from large-scale BIBSEM mosaics and from Optical Light Microscopy (OLM)
images can be combined for multi-scale analysis, ranging from 20 nm to cmscale.
ƒ
Because the Pore-size Density Distribution (PsDD) follows a power law, the
amount of micropores can be extrapolated from OLM imagery.
ƒ
The Total Pore Density (TPD) has a dominant control on electrical resistivity but
the rock type has to be taken into account because pore network architecture
controls the connectivity.
ƒ
Pore network connectivity can be predicted from a combination of pore shape
and spatial analysis parameters (NNCF - Nearest Neighbor Connectivity Factor).
SIGNIFICANCE
Verwer et al. (2011) postulate that electrical resistivity and Archie’s cementation factor
m are directly related to the number of pores and pore throats. This hypothesis is based
on Digital Image Analysis (DIA) of thin sections with Optical Light Microscopy (OLM)
that has a limited resolution of 6-7 μm/pixel. Subsequent tests by analyzing highresolution Micro-CT scans and relating pore throat distributions from MICP
measurements to electrical resistivity corroborated Verwer’s findings (Norbisrath, 2011).
However, finite element resistivity modeling performed on the μ-CT tomograms
indicates the importance of the micropores on electrical properties, which lie below the
resolution of the Micro-CT method.
Figure 1. Zoom into Broad-Ion-Beam (BIB) cross-sectioned and investigated area (red outline) of sample
22. Cut-out on right shows virtually a two-dimensional surface with sub-micron-scale pores ready for
subsequent segmentation and quantification.
83
To assess the influence of the micropores in a quantitative manner, we use a new
method that combines Broad-Ion-Beam milling (Desbois, 2011) and subsequent SEM
image mosaic acquisition (BIBSEM). With this method the sub-micron architecture of
the rock becomes quantifiable and the DIA can be extended from millimeter down to
nanometer scale (Figure 1). This multi-scale quantification is particularly necessary in
carbonate rocks, which are heterogeneous across several length scales.
METHOD AND DATASET
The cutting-edge new BIBSEM technique was performed at the RWTH Aachen in a
joint collaboration. The method utilizes a JEOL SM-09010 cross section polisher to
produce nanometer-precision flat surfaces by milling the rock down with an argon ion
beam. The BIB-milling step is necessary because quantification of nanometer-scale pores
requires large nanometer-precision flat surfaces. BIB milling produces surfaces of up to
2 mm2 total area, in contrast to Focused Ion Beam milling (FIB), which often introduces
surface damage and where investigated areas are 100 times smaller. The large BIB
surfaces are investigated at 5000x and 15000x magnification (resolution: 58.6 nm/pixel
and 18.5 nm/pixel, respectively). The acquired BIBSEM mosaics are composed of around
500 images each, covering up to 1 mm2. Segmentation into pore and solid phase then
allows for Digital Image Analysis. Results from BIBSEM-DIA are combined with DIA
from OLM for multi-scale analysis of pore geometry, and detected pore sizes span from
tens of nanometers to tens of millimeters.
Four samples were chosen from different depositional and diagenetic environments in
order to compare their distinct microstructures. Two of the samples have previously
been imaged in 3D with Micro-CT scanning. All of the samples have been measured for
their electrical resistivity with a NER AutoLab 1000 system, analyzed with Mercury
Injection Capillary Pressure (MICP) methods, and investigated on their macropore
structure with DIA from OLM on epoxy impregnated thin sections. The four rock
samples are: 1)Oolitic Grainstone, 2) Wackestone, 3) Travertine and 4) Dolomite. All the
samples have a similar porosity of around Ø = 16% to minimize the effect of differing
porosity but to allow the assessment of controls of the pore geometry on electrical
resistivity, i.e., the cementation factor m.
Pore Size Density Distribution (PsDD)
PsDDs are analyzed by distributing the pores into logarithmically spaced bins
according to their areas. Rather than their size equivalent diameter the total crosssectional areas are used because electrical current uses the entire area of the opening to
flow, unlike fluids which are more dependent on the shape of the opening due to
capillary forces. The PsDDs are then normalized by total area and bin width and the
resulting normalized pore densities are plotted on log scale (Figure 4).
Pore Network Connectivity (NNCF)
The Nearest Neighbor Connectivity Factor (NNCF) gives an indication of the
connectivity of the pore network. The hypothesis is that the closer the next pore, the
more likely a connection exists. The NNCF is calculated by relating the perimeter of the
pore to the distance to the nearest neighboring pore (NND). This indirect approach has
to be taken as 3D pore network connectivity is very hard to assess from 2-Dimensional
imagery. Finally the values for all the pores in each sample are averaged (Table 1; Figure
5, left).
84
Figure 2. Overviews of the BIB-cross-sectioned areas of the 4 samples with accompanying high-resolution
images at 15000x magnification.
RESULTS
The large BIBSEM image mosaics reveal the diverse microarchitectures of the different
rock types. Around 80-90% of the detected pores were invisible with previous imaging
techniques. Binarized into solid phase and pore phase the mosaics illustrate the
micropore structure of the rock (Figure 3).
Table 1. Petrophysical parameters for the 4 analyzed samples, including the cementation factor m, Helium
Porosity, the Nearest Neighbor Connectivity Factor (NNCF), the Total Pore Density (TPD) for both
OLM+5kx and OLM+15kx magnification imagery, and the slope of the regression line of the Pore Size
Density Distributions (PsDD) from Figure 4.
Sample
m
22
49b
ST36A
WR56.15
2.61
1.72
3.36
2.20
Porosity
[%]
16.2
15.9
14.0
16.8
NNCF
2.09
1.85
1.06
1.40
TPD 5kx
[pores/mm²]
82786
134042
25800
14351
TPD 15kx
[pores/mm²]
222450
414225
222671
44228
Slope
-1.66
-1.80
-1.69
-1.53
85
Total Pore Density (TPD)
The amount of pores per square millimeter varies drastically depending on imaging
resolution. Pore count per square millimeter is 3 to 8 times higher when analyzed with
15000x magnification instead of 5ooox magnification (Table 1). At all resolutions, Total
Pore Density (TPD) is highest in the Wackestone (22) and lowest in the Dolomite
(WR56.15). This is also directly evident from the binarized mosaics (Figure 3). Not
directly visible, however, is that TPD at high magnifications is equally high (around
222000 pores/mm²) in the Travertine (ST36A) as in the Wackestone (49b).
Figure 3. Binarized microstructures of the 4 different rock samples at 5000x magnification. Cementation
factor m, Total Pore Density (TPD), and Nearest Neighbor Connectivity Factor (NNCF) are shown for each
sample. Scale bar is 100 μm.
(Micro-) Pore Structures of Different Carbonate Rock Types
The Ooid Grainstone (sample 22) has a well-connected pore network between
isopachous, bladed cements, covering the partially microporous ooid grains. This pore
network is considered to allow high electrical conductivity, but pore count is low and
tortuosity is increased because the interior of the large ooid grains is not very well
connected. As a result the ooid grainstone has a high cementation factor (m = 2.61).
The Wackestone (sample 49b) is made up almost entirely of a matrix of homogeneous
microspar (microspar crystals: ~1 μm diameter) with very little shell fragments or larger
pores. These small crystals form a very well-connected pore network with low tortuosity
and a very high pore density, which results in the low measured cementation factor (m =
1.72).
86
The Travertine (sample ST36A) has a very scattered and diverse porosity network.
Porous areas are isolated by very dense areas without visible porosity. This, together with
the low pore count, results in the highest observed cementation factor (m =3.4).
The Crystalline Dolomite (sample WR56.15) is characterized by a wide and angular
intercrystalline pore network, which is similar in structure to the Wackestone’s pore
network, just at a larger scale (dolomite rhombs: ~100 μm diameter). The decreased
pore density, however, makes conditions less favorable for the conduction of electrical
charge, which results in a higher cementation factor (m = 2.2).
Figure 4. Pore-size Density Distributions (PsDD). Log of pore densities in exponentially growing pore
size bins, colors represent different imaging techniques and resolutions. Data exhibits a linear behavior,
i.e. predictability at each resolution with high R² values. The regression line is steeper for the more
microporous samples. Pore areas are expressed in size equivalent diameters for better
comprehensibility.
Pore Size Density Distribution (PSDD)
The most arresting finding from this multi-scale pore-structural investigation is that
the PSDDs show a linear behavior at all scales and resolutions (Figure 4) when plotted
on log-log paper. The slope of the regression line through the combined PsDDs is
steepest for the sample with the highest TPD (49b – Wackestone) and lowest for the one
with the lowest TPD (WR56.15 – Dolomite).
87
INTERPRETATION AND IMPLICATIONS
(Micro-) Pore Structures and Pore Network Connectivity (NNCF)
Visual analysis of micro-architectures from BIBSEM mosaics of different rock types can
qualitatively describe their pore network connectivity (Figure 3). It is evident from the
mosaics that the Ooid Grainstone (sample 22) has the best connected pore network and
that the Travertine (sample ST36A) has the least inter-connected pore structure. Visual
comparison between the two crystalline pore structures of the Wackestone (sample 49b)
and the Dolomite (sample WR56.15) is more difficult from the binarized overview images
(Figure 3), because the excellent connectivity of the Wackestone only becomes evident at
higher resolutions (Figure 2; top right).
By using spatial analysis software (GIS), this visual comparison can be enhanced and
the connectivity can be quantified. The NNCF can separate rocks with well-connected
pore networks (e.g. sample 22 - Ooid Grainstone – high NNCF = 2.09) from those which
are less well connected (e.g. sample ST36A – Travertine – low NNCF = 1.06), where TPD
are the same.
A combination of TPDs with NNCFs can give a good estimate of the electrical flow
properties of a pore network (Figure 5).
Figure 5. Cementation factor plotted against NNCF (left) and TPD (right). A correlation is visible, but it is
not yet statistically relevant due to the so far limited dataset.
Electrical transport properties, however, are not directly linked to fluid transport
properties, because electrical current (electrons) can still flow where surface tensional
effects hinders large water molecules to flow (diameter of water molecule is about 0.3
nm). For example, the Wackestone sample shows a very low cementation factor, but the
gas permeability is negligible. This low permeability is caused by the very narrow and
intricate pore network of the sample.
Pore Size Density Distributions (PSDD)
The linear behavior of the PSDDs is the most interesting finding from the new multiscale analysis. This implies that TPD can be estimated from analysis of the pore network
at a limited range of resolution, i.e. a single imaging technique (OLM or BIBSEM). In
combination with pore throat size distributions from MICP measurements, one could
calculate the pore cavity to pore throat size ratio, which is a very important factor for
electrical transport properties, as big parts of larger pores are "dead volume" behind
narrow pore throats. However, when reaching the limit of the resolution at around 20
nm (at 15000x magnification) and 60 nm (at 5000x magnification), the pore size
information becomes less reliable. This is evident from a kink at the upper end of the
Pore-size Density Distributions (Figure 4).
88
89
Figure 6. Multi-scale insight into Ooid Grainstone pore structure. Combination of OLM image (A) from blue-epoxy impregnated thin-section and BIBSEM
mosaics at 5000x (B) and 15000x (C) magnifications illustrates heterogeneity at different magnifications.
Total Pore Density (TPD)
Pore count statistics from OLM and BIBSEM still show a general trend of increased
electrical conductivity (lower m) with increased pore density (Table 1; Figure 1).
However, the effect of pore network architecture on the different rock types has to be
considered. The dolomite sample WR56.15 shows comparably less pores, but the
excellent connectivity (higher NNCF) and low tortuosity of the crystalline microstructure
results in a moderate cementation factor (Table 1 or Figure 5).The increased TPD at
higher resolution illustrates how visible porosity and pore detection are resolutiondependent. The higher the resolution, the more pores you resolve. The question remains,
up to which resolution will you find more pores (i.e. what is the physical lower limit of
pore sizes)? It is also necesary to determine an endpoint when extrapolating amounts of
(nano-)pores from the linearly distributed PsDDs.
Implications
To understand and predict the effect of the pore structure on the electrical behavior in
carbonate rocks, it is paramount to analyze both the macro- and microstructure of their
pore system (Figure 6). This is a consequence of their multi-scale heterogeneity.
Nevertheless, this study indicates that pore size distributions can be predicted from DIA
from a single image source (i.e. at a single imaging resolution), as they follow a power
law. The amount of pores (pore density) at macro- and microscales has proven to be an
essential factor when predicting electrical resistivity; hence this finding could become
useful for reservoir characterization. However, it has to be tested on a larger dataset, one
that consists of a broader range of rock types and also different amounts of porosity.
REFERENCES
Desbois, G., Urai, J. L., et al., 2011, High-resolution 3D fabric and porosity model in a tight gas
sandstone reservoir: A new approach to investigate microstructures from mm- to nm-scale
combining argon beam cross-sectioning and SEM imaging: Journal of Petroleum Science and
Engineering, v. 78, p. 243-257.
Norbisrath, J. H., Eberli, G. P., Weger, R. J., Knackstedt, M., and K. Verwer, 2011, Modeling
electrical resistivity in carbonates using micro-CT scans and assessing the influence of
microporosity using MICP: CSL Annual Review Meeting.
Verwer, K., Eberli, G. P., and R. J. Weger, 2011, Effect of pore structure on electrical resistivity in
carbonates: AAPG Bulletin, v. 95, p. 175-190.
Weger, R. J., Eberli, G. P., Baechle, G. T., Massaferro, J. L., and Y.-F. Sun, 2009, Quantification of
pore structure and its effect on sonic velocity and permeability in carbonates: AAPG Bulletin,
v. 93, no. 10, p. 1297-1317.
90
USING CLUMPED ISOTOPES TO UNDERSTAND
EARLY DIAGENESIS
Peter K. Swart, Monica M. Arienzo, Sean T. Murray,
Yula Hernawati, James S. Klaus, and Donald F. McNeill
KEY FINDINGS
ƒ
Using the clumped isotope thermometer in combination with conventional stable
oxygen isotopic analysis we have determined the G18O of fluid involved in the
formation of corals, bird eggs, and early diagenetic carbonates.
ƒ
In every instance in which the G18O of the depositional fluid was calculated, the
values reflected the composition of the fluids which were actually present.
ƒ
The use of clumped isotopes will revolutionize the interpretation of diagenetic
process in carbonate rocks.
SIGNIFICANCE
Several seminal papers have established patterns in the behavior of geochemical tracers
during early diagenesis (Allan and Matthews, 1982; Lohmann, 1987). These include the
identification of enrichments in the G18O at sub-aerial exposure surfaces, the inverted-J
isotopic signal, the relatively constant G18O values within vadose and freshwater phreatic
zones, a gradual enrichment within the mixing zone, and heavy G18O values within the
marine phreatic zone. All of these signatures are predicated on the basis of a constant
temperature with the G18O reflecting a change in the source or evaporation history of the
water. The use of the clumped isotopic measurements allows an assessment of the
temperature and therefore using the conventional G18O measurement the true G18O of the
water can be measured. This knowledge can lead to substantially improved
interpretation of the paragenetic sequence.
STUDY AREAS
In order to assess the potential for clumped isotopes to reveal information on the G18O
of the fluids involved in precipitation of carbonates we have chosen materials from
modern, Holocene, and Pleistocene carbonates. In each of these cases we have measured
the '47 to determine the temperature and then used this temperature in conjunction
with known temperature-G18O calibrations to calculate the G18O of the precipitating fluid.
The samples include the following.
Modern Corals: Modern corals have been chosen from areas such as Tobago which are
known to be influenced by riverine sources of water (Orinoco) and also corals which
have not experienced any freshwater influence (coastal reefs from Florida). These
different areas should have different G18O of the waters.
Bird Eggs: The G18O of bird eggs have been suggested to be related to the G18O of the
water in which the birds forage. We have compared the clumped isotope signature of
91
eggs from the same species (Great Egrets) living in hydrologically distinct regimes as
well as eggs from Black Skimmers and Cormorants living in northern latitudes which
have isotopically more depleted G18O values.
Clumped Isotopic Signature in Characteristic Diagenetic Zones: The inverted J signal is
a commonly seen signal in early diagenetic carbonates. It is believed to result from the
alteration of the marine sediments by a large pool of meteoric water which imparts its
G18O upon the carbonate leading to the formation of a meteoric carbonate line (MCL).
The materials used for determining these trends are taken from cores drilled during the
Bahamas Drilling Project and the Dominican Republic Drilling Project.
RESULTS
Modern Corals: The G18O of the water calculated using the Ghosh et al. equation (Ghosh
et al., 2006) equation and the Leder et al. equation (Leder et al., 1996) yield values close
to modern seawater for
corals collected from Florida.
Corals which have been
collected
from
regions
clearly
influenced
by
freshwater have much more
depleted G18O values than
than corals living under
normal marine conditions.
Bird Eggs: The calculated
temperatures of the shells
from the Everglades are
within the expected error
(40-44oC), while the birds
from NY and CT have lower
temperatures (Figure 1).
Specimens collected from
the
Everglades
had
18
calculated G O values of the
water
which
were
significantly more positive
than samples from Florida
Bay. In contrast samples of
Black
Skimmers
and
Cormorants, from New York
and
Connecticut
respectively,
had
more Figure 1. Upper panel shows the G13C and G18O of eggshells from three
of birds, 2 specimens of Great Egrets from South Florida
depleted calculated G18O species
(Everglades and Florida Bay), a Black Skimmer (NY) and a
water values. The calculated Cormorant (CT). The calculated temperatures of the shells from the
water compositions are in Everglades are within the expected error (40-44oC), while the birds
agreement
with
values from NY and CT have lower temperatures. The calculated water
measured
in
the compositions are in agreement with values measured in the
environment from which they were collected.
environment from which
they were collected.
92
Clumped Isotopic signatures associated with the Inverted J: The inverted J pattern has
been described as the pathway by which carbonate samples are altered when exposed to
a large volume of meteoric fluid with a constant G18O value (Lohmann, 1987). On a plot
describing G13C and G18O, the pathway of evolution describes an inverted J pattern
forming a meteoric calcite line (MCL) which is representative of calcite in equilibrium
with the meteoric fluid at the temperature of diagenesis (Figure 2). Within a singal
geological deposit multiple MCLs can be recognized representing alteration by a variety
of fluids at different
times.
Analyses of
such samples from a
meteorically altered
deposit
in
the
Dominican Republic
not only confirms the
notion
that
the
different
MCLs
represent alteration in
different fluids, but
also that they have
been
altered
in
different
temperatures (Figure
3). It is suggested
that this alteration
occurred
at
progressively
lower
stands of sea level,
with the heaviest and
Figure 2. G13C and G18O values from Pleistocene cores from the Dominican
coldest MCL having drilling project. Trends describe three different MCL lines. The green box
taken place at the
represents approximate composition of the original sediment.
lowest sea-level stand.
INTERPRETATION AND IMPACT
The data presented in this abstract clearly show the potential of using clumped isotopes
in conjunction with conventional stable isotope analyses. In all instances in which
modern carbonates were formed in environments with elevated G18O values, the
calculated G18O of the fluids were also elevated. In instances where the G18O were
supposedly depleted, the calculated G18O values were also depleted.
93
Figure 3. Temperature and water G18O calculated using clumped isotopes for samples shown in Figure 2.
Each MCL defines a separate trend with the most depleted G18O values representing alteration at
progressively lower temperatures and probably lower sea levels. The most positive values represent the
original G18O of the seawater in which the carbonates were formed.
REFERENCES
Allan, J. R., and R. K. Matthews, 1982, Isotope signatures associated with early meteoric
diagenesis: Sedimentology, v. 29, p. 797-817.
Ghosh, P., Adkins, J., Affek, H., Balta, B., Guo, W. F., Schauble, E. A., Schrag, D., and J. M. Eller,
2006, C-13-O-18 bonds in carbonate minerals: a new kind of paleothermometer: Geochimica
et Cosmochimica Acta, v. 70, p. 1439-1456.
Leder, J. J., Swart, P. K., Szmant, A. M., and R. E. Dodge, 1996, The origin of variations in the
isotopic record of scleractinian corals: I. Oxygen: Geochimica et Cosmochimica Acta, v. 60, p.
2857-2870.
Lohmann, K. C., 1987, Geochemical patterns of meteoric diagenetic systems and their application
to the study of paleokarst, in James, N. P., and P. Choquette, Eds., Paleokarst: Berlin, Springer
-Verlag, p. 58-80.
94
NEW INSIGHTS INTO DOLOMITIZATION USING
CLUMPED ISOTOPES
Sean T. Murray, Monica M. Arienzo, and Peter K. Swart
KEY FINDINGS
ƒ
Through the measurement of the '47 of dolomites, it has become possible to
constrain the temperature and oxygen isotopic value of the fluid involved in the
formation of dolomite. The oxygen isotopic composition in turn relates to the
salinity of the fluid.
ƒ
The analysis of dolomites from locations in which the temperature and
environment of dolomitization is reasonably well constrained allows us to
determine the most suitable equation with which to back calculate the G18O of the
fluid involved in dolomitization.
SIGNIFICANCE
The ever present “Dolomite Problem” has made the understanding of dolomite systems
difficult in the past (Land, 1980). With the advent of clumped isotopes (Ghosh et al.,
2006), it has become possible to approach questions on the formation of dolomites from
a new angle that is not confounded by unknowns such the G18O of the fluid. Despite the
promise of an independent thermometer provided by clumped isotopes, old problems
regarding which of the five equations link temperature to the G18O of the fluid are still
present. However, by utilizing young dolomites that formed in an environment that is
reasonably well constrained, this study is able to make a best guess at which of these
equation is most suitable.
STUDY AREAS
In this study we utilize samples from two areas, The Bahamas, a locality with a
reasonably well constrained diagenetic history and an older dolomite of Carboniferous
age with a more complicated history. The younger rocks are derived from a 168m deepcore drilled on the island of San Salvador in the Bahamas and extensively studied
(Dawans and Swart 1988; Supko 1977; Swart et al., 1987). This core displays an
extended dolomite section replacing middle Miocene to late Pliocene carbonates. These
dolomites are texturally mature but formed as recently as 150,000 yr BP. Through
extensive isotopic and petrographic studies, it has been determined that these dolomites
were formed in two phases by seawater with near normal composition.
The older dolomites are derived from the Mississippian aged Madison Formation at
Sheep Mountain in Wyoming. Dolomites from this locality are fine-crystalline and have
been interpreted as forming during transgressive sea-level cycle changes associated with
the reflux of hypersaline brines from evaporitic lagoons (Smith et al., 2004). This
interpretation was supported by the existence of solution collapse breccias which are
associated with the existence of evaporitic lagoons (Sonnenfeld, 1996). The dolomites are
then believed to have been altered by meteoric waters which reset some of the dolomites
isotopically (Moore, 2001).
95
The data from these two localities are then compared with the Latemar, in the Italian
Alps formed during the middle Triassic. The dolomites at this location show varying
degrees of textural maturity with the most mature in the center of the buildup, but there
is a sharp transitional boundary with the out-lying calcites. The Latemar dolomites form
a mushroom shaped cap approximately 2.5 km across and 400 m high which cross cuts
multiple formations and depositional features. Their formation has been suggested to be
either unmodified, hydrothermal sea water driven by plutonic activity (Wilson et al.,
1990) or modern diffuse flow fluids derived from the mid-ocean ridges (Carmichael and
Ferry, 2008; Carmichael et al., 2008). These dolomites were the subject of a recent
paper utilizing clumped isotopes by Ferry et al. (2011).
RESULTS
Figure 1. (A) The lines represent the interpretation using
the dolomite-temperature line from Vasconcelos (2005).
The San Salvador data plots as forming from a fluid with
a G18O between +2‰ to +4‰. This is too high and
inconsistent with previous interpretations.
Figure 1. (B) These are the same data, but adjusted using
the constraints of Shepard and Schwarz (1970) which puts
the San Salvador data in range consistent with formation
from normal marine waters.
96
Measurements using the
clumped isotope method were
made on a section of the San
Salvador core spanning 54m
to 67m depth and on the
Sheep Mountain samples
from 299m to 317m. The
temperatures were calculated
using the equations of Ghosh
et al. (2006) and corrected
using the methods of Dennis
et al. (2011).
The samples
from San Salvador have a
calculated temperature of
formation of 31 to 39°C with
an average standard error of
±1.7°C. The temperatures
show a slightly increasing arc
down core.
Samples
from
Sheep
Mountain
display
a
temperature range of 40 to
62°C
with
an
average
standard error of ±1.5°C. The
temperatures trend towards
higher temperatures with
increased
amounts
of
dolomite present in the
sample.
Although
the
clumped
isotope method provides
temperature, the calculation
of
the
fluid
isotopic
composition necessitates the
use of one of the five
equations (Fritz and Smith,
1970; Northrop and Clayton,
1966; O'Neil and Epstein, 1966 ; Sheppard and Schwarcz, 1970; Vasconcelos et al., 2005)
which link the temperature of formation and the į18Owater with temperature. This
problem has been recognized as a major stumbling block in dolomite interpretation over
the past 30 years (Land, 1980). In order to determine which is the most appropriate
equation we used the dolomites from San Salvador which are known to have formed at
near surface temperatures and from fluids which were near seawater or maybe slightly
elevated in their į18Owater values. The most recent equation (Vasconcelos et al. 2005) of
the five provides fluid compositions greater than+4‰. We consider these values too
positive to be realistic. The other extreme, the equation of Northrop and Clayton (1966),
gives water values which are -4‰. The equation of Sheppard and Schwarz (1970)
provides fluid values which are in closest agreement with the previous interpretations
(Figure 1).
INTERPRETATION AND IMPACT
San Salvador: The dolomitized interval from San Salvador represents an alteration from
a hard crystalline dolomite showing mimetic replacement of the precursor to a
Figure 2. The alternation (LH) between the different types of dolomite described in the text used to select
the samples for temperature measurements (figure from Dawans and Swart, 1988). Temperatures (RH)
increase into the middle of the alternation and then decrease towards the base. The high temperature at
the base represents the start of the next alternation. See Figure 3 for į18Owater values.
97
microsucrosic dolomite. These alterations occurred repeatedly throughout the core and
each one was associated with a sub-aerial exposure surface and a change in the
stoichiometry, Sr content, and G18O (Figure 2).
Within this one alternation of
dolomite types, the measured
temperature ranged from 31 to 29oC
with higher values being prevalent
in the central portion of the
alternation (Figure 2). The change
in
temperature
is
positively
correlated with the G18O of the
fluids so that the warmer fluids
represent fluids elevated in 18O
(Figure 3).
The range of
temperatures measured in the San
Salvador dolomites seem too high
by 10-15oC, perhaps a result of the
Figure 3. Correlation between temperature and fluid
fact that the Ghosh et al., (2006)
G18Owater from San Salvador.
equation may not be directly
applicable to dolomites. It should
be noted that in the previous study on dolomites (Ferry et al., 2011) a completely
theoretical approach (Guo et al., 2009) was used which tends to yield slightly lower
temperatures than those presented here. Lower temperature would in turn lead to lower
calculated fluid values.
Sheep Mountain: Temperatures
calculated for the Sheep Mountain
dolomites show higher values (40 62oC) than those from San Salvador.
These higher values could reflect
partial resetting of originally lower
formation
temperatures
during
burial. As these rocks are not 100%
dolomites (in contrast to San
Salvador), we have estimated the
G18O of the fluid using a combination
of the calcite-water (Kim and O'Neil,
1997) and dolomite-water equations
(Sheppard and Schwarcz, 1970).
These rocks show a range of values
from ~0‰ to +4‰ (excluding 2
points which have significantly lower
G18O values) with a tendency for
values to be the most positive
immediately below the breccia
(Figure 4).
This would tend to
support the hypothesis that the
dolomitization is associated with the
collapsed breccia as previously
proposed.
98
Figure 4. The G18O of the fluids from Sheep
Mountain. The collapsed breccia is situated around
430 m and so values become more enriched towards
the feature.
Latemar: There has been only one previously published article on the application of
clumped isotopes to the study of dolomitization (Ferry et al., 2011). These workers
measured a temperature range of between 50 to 70 °C with the į18Owater averaging -0.3 ±
1.7 ‰. This is in stark contrast to a previous study which measured fluid inclusion
temperatures above 150oC (Wilson et al., 1990). The later study concluded that the fluid
inclusions had been reset by a later event. It was suggested that the temperature range
of 50-70oC reflected a partial resetting of the temperature during later burial. It is also
important to note that this study used the Vasconcelos et al. (2005) equation to calculate
the į18Owater values. Based on our work and considering the nature of the samples used
to construct that relationship we feel the Sheppard and Schwarcz (1970) equation may be
more appropriate. Changing the equation would produce diagenetic fluids with lower
į18Owater values than originally suggested.
CONCLUSIONS
Although the application of clumped isotopes has the potential to revolutionize the
study of dolomitization, there are still some uncertainties. These include:
1) Which is the correct temperature-water relationship to use when calculating the G18O
of the dolomitizing fluid?
2) Is the Ghosh et al. (2006) equation applicable to dolomite?
3) To what degree does the original temperature signature imparted during the
formation of the dolomite, become altered during burial?
REFERENCES
Carmichael, S. K., and J. M. Ferry, 2008, Formation of replacement dolomite in the Latemar
carbonate buildup, Dolomite, northern Italy: Part 2. Origin of the dolomitizing fluid and the
amount and duration of fluid flow: American Journal of Science, v. 208, p. 885-904.
Carmichael, S. K., Ferry, J. M., and W. F. McDonough, 2008, Formation of replacement dolomite
in the Latemar carbonate buildup, Dolomites, northern Italy Part 1. Field relations,
mineralogy, and geochemistry: American Journal of Science, v. 308, p. 851-884.
Dawans, J., and P. K. Swart, 1988, Textural and geochemical alternations in late Cenozoic
Bahamian dolomites: Sedimentology, v. 35, p. 385-403.
Dennis, K. J., Affek, H. P., Passey, B. H., Schrag, D. P., and J. M. Eiler, 2011, Defining an absolute
reference frame for 'clumped' isotope studies of CO2: Geochimica et Cosmochimica Acta, v. 75,
p. 7117-7131.
Ferry, J. M., Passey, B. H., Vasconcelos, C., and J. M. Eiler, 2011, Formation of dolomite at 40-80
degrees C in the Latemar carbonate buildup, Dolomites, Italy, from clumped isotope
thermometry: Geology, v. 39, p. 571-574.
Fritz, P., and D. G. W. Smith, 1970, The isotopic composition of secondary dolomites: GCA, v. 34,
p. 1161-1173.
Ghosh, P., Adkins, J., Affek, H., Balta, B., Guo, W. F., Schauble, E. A., Schrag, D., and J. M. Eller,
2006, C-13-O-18 bonds in carbonate minerals: A new kind of paleothermometer: Geochimica
et Cosmochimica Acta, v. 70, p. 1439-1456.
Guo, W., Mosenfelder, J. L., Goddard, W. A., III, and J. M. Eiler, 2009, Isotopic fractionations
associated with phosphoric acid digestion of carbonate minerals: insights from the first
99
principles theoretical modeling and clumped isotope measurements: Geochimica et
Cosmochimica Acta, v. 73, p. 7203-7225.
Kim, S. T., and J. R. O'Neil, 1997, Equilibrium and nonequilibrium oxygen isotope effects in
synthetic carbonates: Geochimica et Cosmochimica Acta, v. 61, p. 3461-3475.
Land, L. S., 1980, The isotopic and trace element geochemistry of dolomite: the state of the art:
SEPM Special publication 28, p. 87-110.
Moore, C., 2001, Carbonate reservoirs: Porosity evolution and diagenesis in a sequence
stratigraphic framework: Developments in Sedimentology, v. 55: Amstterdam, Elsevier, 444 p.
Northrop, D. A., and R. N. Clayton, 1966, Oxygen isotope fractionation in systems containing
dolomite: J. Geol., v. 74, p. 174.
O'Neil, J. R., and S. Epstein, 1966, Oxygen isotope fractionation in the system dolomite-calcite
carbon dioxide: Science, v. 152, p. 198-201.
Sheppard, S. M. F., and H. P. Schwarcz, 1970, Fractionation of carbon and oxygen isotopes and
magnesium between coexisting calcite and dolomite: Contrib. Mineral Petrol., v. 26, p. 161.
Smith, L. B., Eberli, G. P., and M. Sonnenfeld, 2004, Sequence-stratigraphic and paleogeographic
distribution of reservoir-quality dolomite, Madison Formation, Wyoming and Montana, in
Grammer, G. M., Eberli, G. P., and P. M. Harris, Eds., Intergration of outcrop and modern
analogues in reservoir modeling: American Association of Petroleum Geologists Memoir, p.
94-118.
Sonnenfeld, M. D., 1996, Sequence evolution and hierarchy within the lower Mississippian
Madison Limestone of Wyoming, in Longman, M. W., and M. D. Sonnenfeld, Eds., Paleozoic
Systems of the Rocky Mountain region, Society for Economic Paleontologists and
Mineralogists (Society for Sedimentary Geology) Rocky Mountain Section, p. 165-192.
Supko, P. R., 1977, Subsurface dolomites, San Salvador, Bahamas: Journal of Sedimentary
Petrology, v. 47, p. 1063-1077.
Swart, P. K., Ruiz, J., and C. W. Holmes, 1987, Use of strontium isotopes to constrain the timing
and mode of dolomitization of Upper Cenozoic sediments in a core from San Salvador,
Bahamas: Geology, v. 15, p. 262-265.
Vasconcelos, C., McKenzie, J. A., Warthmann, R., and S. M. Bernasconi, 2005, Calibration of the
G18O paleothermometer for dolomite precipitated in microbial cultures and natural
environments: Geology, v. 33, p. 317-320.
Wilson, E. N., Hardie, L. A., and O. M. Phillips, 1990, Dolomitization front geometry, fluid-flow
patterns, and the origin of massive dolomite - the Triassic Latemar buildup, Northern Italy:
American Journal Of Science, v. 290, p. 741-796.
100
SPELEOTHEMS: A MODEL SYSTEM FOR THE STUDY OF
FLUID INCLUSIONS AND CLUMPED ISOTOPES
Monica M. Arienzo, Sean T. Murray, Hubert B. Vonhof1,
and Peter K. Swart
1)
Vrije Universiteit Amsterdam, Amsterdam, The Netherlands
KEY FINDINGS
ƒ
Oxygen isotopic analysis of fluid inclusions, combined with clumped isotopes,
and the in situ monitoring of calcite precipitation in caves allows for the
determination of the competing influences of temperature and water isotopic
composition in controlling the oxygen isotopic composition of carbonates.
ƒ
The example presented here, a speleothem from the Bahamas, demonstrates that
by applying these various methods we can develop a clear understanding of the
fluids which form these carbonates.
ƒ
The lessons from these examples can be applied to a wide range of diagenetic
carbonates.
SIGNIFICANCE & BACKGROUND
Traditionally carbon and oxygen isotope analyses have been used to unravel the
depositional and diagenetic history of carbonates. However, the į18O of a carbonate is
dependent both upon the variations in temperature as well as the į18O of the water. In
order to solve for the second unknown, an additional proxy is needed which can provide
information on one of the two unknowns. Such proxies might include the ratio of certain
trace elements relative to calcium, fluid inclusions, and/or clumped isotopes.
The motivation for this study is to utilize a speleothem from the Bahamas as a case
study for the application of clumped isotopes and stable isotopic analysis of fluid
inclusions. By applying these various methodologies we hope to gain a better
understanding of the factors which control the oxygen isotopes of the speleothem.
METHODS
This study will focus on a speleothem, sample DC-09, collected from Dan’s Cave on
Abaco Island, Bahamas. Sampling of the speleothem for stable C and O isotopes was
conducted using a computerized micromill and analysis was performed on the Delta Plus
mass spectrometer. Fluid inclusion analysis is the analysis of microscopic, water filled
cavities within the stalagmite. These cavities preserve drip water at the time of formation
and allow for the direct measurement of the G18O composition of the formation water.
Fluid inclusion analyses were conducted at a resolution of about one sample every 1.5
cm. Fluid inclusion isotopes were analyzed at Vrije Universiteit Amsterdam. Analysis
was conducted utilizing the “Amsterdam Device”, an instrument built specifically for the
extraction of water from fluid inclusions (Vonhof et al., 2006). The extracted water was
then measured for oxygen and hydrogen isotopes.
Carbonate clumped isotope analysis is based on the measured abundance of the rare
isotopes in the carbonate, such as the 13C-18O bonds (Ghosh et al., 2006) . The 13C-18O
101
‘clump’ is of interest to geochemists because as temperatures decrease, clumping
increases, independent of the isotopic composition of the formation water . This method
therefore allows temperature to be determined without prior knowledge of the d18O of
the fluid involved in precipitation. The same samples utilised for fluid inclusion analysis
were also measured for clumped isotopes allowing for a direct comparison between the
two proxies. Clumped isotope analyses were conducted at the University of Miami and
have been standardized using the method from Dennis et al. (2011). Carbonate
stalagmite samples have been run in triplicate with an average ¨47 standard error of
0.007 which equates to a 1.5 °C range in temperature. The clumped isotope analyses
have been conducted using the Thermo MAT-253 in the Stable Isotope Laboratory.
RESULTS
Plotting the į18Oc of the calcite with respect to age demonstrates that there are
significant variations in the oxygen isotopic value of the calcite (Figure 1). There is an
observed increase in the į18Oc followed by a very rapid decrease in the į18Oc value. As
discussed above, determining the drivers of these changes is problematic in that they
could either be temperature related or reflect variations in the į18Ow of the fluid.
The fluid inclusion and clumped isotope records support a change in the į18Oc of the
fluid precipitating the stalagmite. The fluid inclusion į18Ow (water) increases in concert
with increases in the į18Oc (Figure 1). This is then followed by a decrease in the į18Ow as
the carbonate oxygen isotopes decrease. Utilizing the į18Oc and į18Ow results,
temperature can be calculated. For stalagmites, this relationship was determined from
cave monitoring experiments by Tremaine et al., 2011. Applying this equation supports
additional temperature variation associated with the changes in the į18Oc and į18Ow
(Figure 2).
Figure 1. Blue line represents the į18Oc record from the calcite. Blue squares are the į18Ow derived from
the fluid inclusions and orange circles are Ʃ47 measurements, note the axis is reversed.
102
The clumped isotope record supports the fluid inclusion data (Figure 1). Temperature
from Figure 2 is calculated using the Ghosh et al. (2006) equation. The į18Ow is
calculated from temperature using the equation from Tremaine et al., (2011). There is an
offset between the Ʃ47 and fluid inclusion į18Ow records and also an offset between the
two temperature records (Figure 2). The calculated į18Ow from the clumped isotope
record is more positive than the fluid inclusion data, with an average offset between the
records of about 0.6 ‰. The temperature offset between the two records is about 9 °C
(Figure 2). This offset is not unique to this speleothem sample. The observed offset is
consistent with other speleothem studies and is thought to be driven by the way the
calcite precipitates in speleothems (Affek et al., 2008; Kluge and Affek, 2012). However,
overall the clumped isotope record supports similar trends to the fluid inclusion results,
with an increase in į18Ow associated with the į18Oc increase and associated temperature
variation.
Although there are still uncertainties regarding the interpretation of the clumped
isotope signal in stalagmites, the į18Ow record from the fluid inclusions is supported by
the calculated oxygen isotope ratio of the water based on the clumped isotopes.
Figure 2. (Left) Blue line with triangles represents į18Ow derived from the fluid inclusions and orange
line with triangles is į18Ow derived from Ʃ47 measurements. (Right) Blue line is the water from fluid
inclusions, orange line is the water calculated from Ʃ47.
FLUID INCLUSION ANALYSIS AT THE UNIVERSITY OF MIAMI
Recent work has been conducted on developing a fluid inclusion extraction device at
the University of Miami. The “Miami Device” is unique because the setup utilizes cavityring down spectroscopy (CRDS) for į18O and į2H analysis of fluid inclusions. The
extraction line at the University of Miami is an in-line system directly interfaced with the
Picarro CRDS isotopic water analyzer and the design of the line is similar to the
103
Amsterdam Device (Vonhof et al., 2006). The extraction line consists of a crusher which
is a modified valve unit, a septum port for the direct injection of water and a water trap
(Figure 3a). The CRDS technique is based on using a near infrared laser to scan over the
H2O spectral range and by measuring the absorption spectra using a ring-down method,
to determine isotopic abundances (Figure 3b). Preliminary results demonstrates a 0.3 NjL
water injection provides ample signal for isotopic analyses with an average standard
deviation of 0.29 ‰ for į18O and 2.2 ‰ for į2H, which is comparable to other fluid
inclusion extraction devices (Figure 3b). This setup will now enable in house isotopic
measurements of fluids from carbonates.
Figure 3. (Top) “Miami Device” on the right with the Picarro CRDS on the left. (Bottom) Screen shot of
the Picarro analyzing 0.3 NjL of water injected into the Miami Device. Top view shows the ppm of H2O,
middle view of the į18O, and bottom view is of the į2H.
104
BAHAMAS CAVE MONITORING
In the summer of 2012, cave monitoring began in Eleuthera, Bahamas. Here, the
temperature of the cave and relative humidity are measured every 2 hours. Additionally,
we are conducting in situ monitoring of calcite precipitation (Figure 4). Calcite is
“farmed” by placing glass slides on top of actively forming stalagmites within the cave
with the slides being collected every 2-3 months (Figure 4). This allows for direct
comparison between the cave environment and the chemistry of the calcite.
The
carbonate is then removed from the glass slide and analyzed for stable C and O isotopes
and clumped isotopes. In Figure 4, preliminary results from clumped isotopes are
shown. The offset between the clumped isotope and actual temperature is 3.4°C , with
the clumped isotopes giving a warmer temperature. As demonstrated, there are still
uncertainties regarding the interpretation of the clumped isotope signal in stalagmites
and by continually monitoring a Bahamas cave and collecting calcite as it forms, a better
understanding of clumped isotopes in speleothems can be developed. Through continual
monitoring, this work may potentially aid in the development of a calibration equation of
Ʃ47 to temperature for speleothems.
Slide 6 - HBC
Figure 4. (Top left) A student descends into the monitoring cave in the Bahamas. Top right: Calcite
farming set up within the cave. (Bottom left) Microscope slide after 2 months in the cave, calcite was
removed from the slide and analyzed for clumped isotopes. (Bottom right) Actual measured temperature
compared with preliminary clumped isotope derived temperature from the calcite farming.
105
INTERPRETATION AND IMPLICATIONS
This case study demonstrates how the application of the clumped isotope and fluid
inclusion methods can aid in understanding the fluid from which carbonate precipitated.
This study demonstrates that changes in the calcite isotopes are a result of changes in the
calcite precipitating fluids, as well as temperature. Combining these two methods may
aid our understanding of diagenesis and of the fluids precipitating carbonates.
Furthermore, future cave calcite farming may provide valuable insight regarding the
processes driving geochemical records.
REFERENCES
Affek, H. P., Bar-Matthews, M., Ayalon, A., Matthews, A., and J. M. Eiler, 2008,
Glacial/interglacial temperature variations in Soreq cave speleothems as recorded by 'clumped
isotope' thermometry: Geochimica et Cosmochimica Acta, v. 72, p. 5351-5360.
Dennis, K. J., Affek, H. P., Passey, B. H., Schrag, D. P., and J. M. Eiler, 2011, Defining an absolute
reference frame for 'clumped' isotope studies of CO2: Geochimica et Cosmochimica Acta, v. 75,
p. 7117-7131.
Ghosh, P., Adkins, J., Affek, H., Balta, B., Guo, W. F., Schauble, E. A., Schrag, D., and J. M. Eller,
2006, C-13-O-18 bonds in carbonate minerals: A new kind of paleothermometer: Geochimica
et Cosmochimica Acta, v. 70, p. 1439-1456.
Kluge, T., and H. P. Affek, 2012, Quantifying kinetic fractionation in Bunker Cave speleothems
using Ʃ47.: Quaternary Science Reivews, v. 49, p. 82-94.
Tremaine, D. M., Froelich, P. N., and Y. Wang, 2011, Speleothem calcite farmed in situ: Modern
calibration of G18O and G13C paleoclimate proxies in a continuously-monitored natural cave
system: Geochimica et Cosmochimica Acta, v. 75, p. 4929-4950.
Vonhof, H. B., van Breukelen, M. R., Postma, O., Rowe, P. J., Atkinson, T. C., and D. Kroon, 2006,
A continuous-flow crushing device for on-line G2H analysis of fluid inclusion water in
speleothems: Rapid Communications in Mass Spectrometry, v. 20, p. 2553-2558.
106
SEISMIC AND GPR IMAGING OF FRACTURES IN
CARBONATE RESERVOIRS USING 3D DIFFRACTION
RESPONSES CAUSED BY FRACTURE INTERSECTIONS
Mark Grasmueck, Tijmen Jan Moser1, and Michael A. Pelissier2
1)
2)
Moser Geophysical Services, Den Haag NL
Marathon Oil Company, Houston USA
KEY FINDINGS
ƒ
Due to kinematic similarity of GPR and seismic wave propagation, GPR data can
be used as a scaled proxy for seismic data to study the use of diffractions for
fracture imaging below the resolution limit of the reflection seismic method.
ƒ
Intersections of perpendicular fracture sets create dihedral and trihedral
scatterers causing recordable diffractions from fracture systems with little or no
vertical displacement and 1/500th of a wavelength fracture width.
ƒ
Fracture intersections and hence the resulting diffractions are direct indicators of
fracture connectivity.
ƒ
Seismic diffraction signal levels are lower than their GPR counterparts. In order
to use diffractions at the reservoir level, special care to preserve and enhance
diffractions must be taken during acquisition and processing.
INTRODUCTION AND SIGNIFICANCE
Typically sub-vertical fracture mapping based on seismic data relies on displacements
of otherwise continuous reflections. Semblance or coherency attributes help make small
displacements visible. In theory, vertical displacements as small as one quarter
wavelength of the highest frequency are resolvable. For a typical carbonate reservoir, a
velocity of 5000 m/s and seismic signal frequency of 50 Hz the smallest visible fracture
displacement of a stratigraphic reflection is 25 m. Such a fracture is a rather large fault.
Fractures with small or zero vertical displacement are thus beyond the resolution of
classical reflection seismic fracture mapping. Borehole imaging and structural modeling
may help estimate the distribution of smaller fractures but are limited by spatial
uncertainty. Besides carbonate reservoirs, tight shale reservoir production and
stimulation would benefit from laterally extensive information about connected fracture
networks with below quarter wavelength displacements.
Diffractions are a promising source for sub-wavelength 3D fracture information.
Diffractions originate from small-scale discontinuities in the subsurface and are
normally treated as noise in conventional seismic processing as they interfere with
continuous reflections. Diffractions already have been successfully used to define oil
bearing karst caverns which previously had not been resolved (Yang et al., 2011).
Similarly, Pomar (2010) showed how diffractions recorded in dense 3D Ground
Penetrating Radar (GPR) data can be used to image complex karst and fracture
networks. Li et al. (2012) propose a geologically plausible model for the origin of karst
diffractions with sub-wavelength spherical or random bodies of slow material (2500–
3200 m/s) embedded in fast carbonate hostrock (6000 m/s). The objective of this paper
107
is to find a geological model for the origin of fracture related diffractions where no karst
voids are present.
HYPOTHESIS AND APPROACH
Our Hypothesis is that seismically recordable diffractions are caused by intersections of
thin fractures with no displacement.
We use high-resolution 3D GPR data as a bridge between synthetic modeling and
seismic reservoir imaging. Natural fracture networks of outcropping reservoir analogs
can be efficiently imaged with 3D GPR and interpreted with the help of the nearby
outcrop (Grasmueck et al., 2012). Due to the kinematic similarity of GPR and seismic
wave propagation, the GPR data can used as a proxy for seismic data to help develop new
diffraction based fracture imaging workflows. Through scaling relationships the GPR
findings are applied to the seismic method and provide guidance for the use of
diffractions for determining previously seismically unresolvable fracture systems at
reservoir depth.
RESULTS
3D GPR Response of a vertical X subhorizontal Fracture Intersection
Figure 1 is a small subvolume of the larger Cassis 3D GPR survey acquired with 5 x 10
cm trace spacing and 200 MHz antennae (Grasmueck et al., 2012). The low amplitude
11º dipping subhorizontal reflections are caused by around 1 mm open joints following
stratigraphic boundaries. When following the reflection bands of these subhorizontal
fractures in timeslices they are lined by small bright spots. The center of the timeslice in
Figure 1 is located on such a bright spot. On unmigrated data the bright spot corresponds
to the apex of a diffraction cone. For a slightly deeper timeslice the spot has the shape of
a circle (Figure 1b). In the 3D migrated data in Figure 1c), d) the diffraction is focused in
a small and elongate high amplitude anomaly. The long axis of the migrated anomaly is
aligned with the intersection line between the two joints (Figure 1 e). Within the entire
3D GPR survey volume hundreds of such small bright spots can be observed. Laterally
extensive fractures are defined by multiple such bright spots aligned in the same fracture
plane.
In the nearby outcrop wall the subhorizontal fractures are continuous over tens of
meters and relatively smooth (Figure 2a). In contrast, the vertical fractures consist of
multiple segments belonging to the same fracture trend. Fracture opening is less than 1
mm, similar to the horizontal fractures. The size of continuous fracture segments is
typically less than 0.5 m, thus smaller than the wavelength of the GPR signal. At the
intersection of vertical and horizontal fractures sharp corner geometries are formed
leading to a blocky appearance of the quarry wall. The dimensions of these blocks are
between 0.1 and 0.5 m (Figure 2b). Within the 3D rock volume the corners form
dihedrals for 2 intersecting fractures or trihedrals (also known as cateye or corner
reflector) when 3 perpendicular fractures intersect. Dihedrals and Trihedrals are known
to be efficient wavefield scatterers. The same geometrical configuration is used in retro
reflectors for the safety of vehicles and back scattering targets in satellite remote sensing
applications. In the case of fractured media the intersection of near perpendicular
fractures creates natural scatterers causing diffractions in GPR and seismic data.
108
Ray-Born Synthetic Modeling of Fracture Intersections
As shown in Figure 2c) a simple cross represents the basic geometrical element of a
fracture intersection. Ray-Born synthetic modeling (Moser, 2012) of crosses with
different sizes (Figure 3a) show that for dimensions of less than one quarter wavelength
point diffraction responses are generated. The larger the cross the higher the amplitude.
The point diffraction does not resolve the individual arms of the cross (Figure 3b). When
exceeding a quarter wavelength, also known as the Rayleigh resolution limit, the
diffraction tails split up as the arms of the cross generate individual diffractions
interfering with each other. As a result the amplitude is concentrated in the apex and
weaker on the split diffraction tails (Figure 3c) also leading to multiple interfering circles
on timeslices as seen for example in Figure 1b.
Figure 1. Subvolume of larger 3D GPR survey
acquired in Cassis Quarry (France). The bright
spot in the center of the top face is caused by the
intersection of a vertical with a sub-horizontal
fracture. a) and b) are unmigrated data. In b) the
top face is 5 cm deeper than in a) showing circular
diffraction pattern. c) and d) are the
corresponding 3D migrated cubes where the
diffraction hyperboloid has been focused into a
small high amplitude anomaly. e) Fracture
Interpretation. Vertical exaggeration of the cube
display is 3.3x.
109
UPSCALING OF DIFFRACTION RESPONSE FROM OUTCROP GPR TO RESERVOIR
DEPTH SEISMIC DATA
While many clear diffraction signatures are observed in dense 3D GPR data, seismic
diffraction examples from reservoir depth are still rare. Table 1 compares the GPR and
seismic parameters relevant for the detection of diffractions in the near surface and at a
typical carbonate reservoir depth of 4 to 5 km. Central frequency of the seismic data is
assumed to be 50 Hz.
The absolute seismic scattering strength for oil- or water-filled fractures is about half of
the air-filled limestone fractures of the Cassis outcrop. The 1 mm fracture opening
observed in outcrop translates into 20 cm fracture width at reservoir level. Dihedrals and
trihedrals at the intersections of fracture segments with extents of less than 25 m cause
point diffractions responses as shown in the modeling results of Figure 3. For proper
sampling of the full diffraction signals seismic surveys need to be acquired with a single
sensor trace spacing of less than 12.5 m to also include the higher than 50 Hz frequency
content of the wavelet. Many recently acquired seismic surveys satisfy this spatial
sampling requirement.
The maximum reflection depth recorded in the Cassis 3D GPR survey is 10 m
corresponding to twenty wavelengths. Using the number of wavelengths for the seismic
cases translates into 1600 m which is too shallow for most reservoirs. The GPR data also
show that clear diffractions are only visible to half or even just a quarter of the maximum
reflection depth. Two factors are responsible for this strong reduction of depth for the
observation of diffractions: 1) Diffraction signals experience spherical spreading
amplitude decay for the down-going and the up-going wavefield. For reflections
spherical spreading affects only the down-going part as the up-going wave can be
approximated by a plane wave. 2) Diffractions depend on the recording of wide
apertures. At a radiation angle of 60º measured from vertical, only half the vertical depth
can be reached with the same signal amplitude. With these amplitude reductions on the
order of one magnitude, seismic diffractions can only be recorded to less than 1000 m
depth. In fact, most published seismic diffraction examples are from seismic data sets
(Berkovitch et al., 2009). In order to overcome this depth limitation and see clear
diffractions at the reservoir level three measures must be taken:
1) Acquire very dense, vertically stacked single sensor data with a sufficiently
high SN ratio. While GPR equipment only has 16 bit dynamic range, the standard 24
bits of seismic AD converters are sufficient to also record weak diffraction signals
originating at reservoir depth.
2) Preserve all diffracted Energy during processing.
3) Separate reflection and diffraction parts of the seismic signal and perform
diffraction analysis on the diffraction only part (Moser and Howard, 2008).
IMPLICATIONS AND CONCLUSIONS
Diffractions are not just caused by karstic voids and caverns. Our hypothesis that
diffractions originate at the intersection of thin fractures with no displacement is
supported by high-resolution 3D GPR data compared to the nearby outcrop and
confirmed by synthetic modeling. Quarter wavelength or smaller dihedrals, trihedrals
created at the intersections of perpendicular fracture sets are efficient point scatterers
with omnidirectional radiation patterns. The non-random distribution of these subRayleigh size discontinuities is caused by fracture trends and patterns providing
information about fracture spacing and fracture continuity of fractured domains. By
110
their nature fracture intersections and hence the resulting diffractions are direct
indicators of fracture connectivity. Seismic diffraction signals from 4-5 km deep
carbonate reservoirs are by an order of magnitude weaker than reflections from the same
depth. To fully harness diffractions and their information about reservoir fracture
systems seismic data need to be acquired densely with high signal-to-noise ratio coupled
with processing optimized for diffraction signal preservation and separation.
Figure 2. Quarry wall below the site where the 3D GPR of Figure 1 were acquired. Subhorizontal fractures
are intersected by vertical fracture segments and form dihedral scatterers.
Figure 3. Ray-Born synthetic data of line crosses with different sizes. Until quarter wavelength cross
diameter point diffraction responses with increasing amplitudes are generated.
111
Table 1. GPR vs. Seismic Scale Comparison. To reach sufficient seismic diffraction strength at reservoir
level an order of magnitude signal dynamic range gain is needed.
REFERENCES
Berkovitch, A., Belfer, I., Hassin Y., and E. Landa, 2009, Diffraction imaging by multifocusing:
Geophysics, v. 74, WCA75–WCA81. doi:10.1190/1.3198210.
Grasmueck, M., Coll, M., Eberli, G. P., and K. Pomar, 2012, Diffraction imaging of sub-vertical
fractures and karst with full-resolution 3D GPR: Accepted for publication in Geophysical
Prospecting.
Li, F., Di, B., Wei, J., and X. Li, 2012, Volume estimation of the carbonate fracture-cavern
reservoir - A physical model study: 74th EAGE Conference & Exhibition.
Moser, T. J., and C. B. Howard, 2008, Diffraction imaging in depth: Geophysical Prospecting, v.
56, p. 627–641. doi: 10.1111/j.1365-2478.2007.00718.x.
Moser, T. J., 2012, Review of ray-Born modeling for migration and diffraction analysis: Studia
Geophysica et Geodætica, In Press.
Pomar, K., 2010, Visualization and quantification of fractures and karst in Cretaceous carbonates,
Cassis, France: University of Miami M.S. Thesis, Open Access Theses, Paper 76.
http://scholarlyrepository.miami.edu/oa_theses/76
Yang, P., Liu, Y. L., Li, H. Y., Dan, G. J., An, H. T, and Y. M. Shao, 2011, Fractured-vuggy reservoir
characterization of carbonate, Tarim Basin, Northwest China: 73rd EAGE Conference &
Exhibition.
112
4D GPR FOR CHARACTERIZATION OF FLUID FLOW IN
CARBONATES: INSIGHTS FROM STRUCTURAL- VS.
STRATIGRAPHIC-CONTROLLED DOMAINS AND
COMPARISON WITH ECLIPSE DYNAMIC MODELING
Pierpaolo Marchesini, Mark Grasmueck, Gregor P. Eberli, and
Ralf J. Weger
KEY FINDINGS
ƒ
4D GPR monitoring of gravity flow in reservoir analogues allows the
characterization of fluid dynamics at the 1-10 m scale.
ƒ
Small-scale stratigraphic boundaries control fluid flow in high-porosity, nonfractured carbonates.
ƒ
Fluid migration in fractured domains is governed by structural discontinuities
such as faults and deformation bands.
ƒ
Deformation bands are responsible for fluid compartmentalization in highlysaturated conditions.
ƒ
Standard approach in dynamic modeling fails to capture the influence of smallscale heterogeneities on fluid dynamics.
INTRODUCTION AND SIGNIFICANCE
Lateral variability makes the characterization of fluid dynamics in carbonate reservoirs
an open challenge. Even a precise reconstruction of stratigraphic boundaries and
fracture networks does not guarantee a comprehensive knowledge of preferential flow
paths. Current characterization of hydraulic parameters largely relies on 0.01-0.1 m scale
laboratory experiments, sample plug measurements, and modeling. However, a large
degree of upscaling prevents these methods from fully reproducing realistic flow
conditions. In this study we used time-lapse 3D GPR (4D GPR) to non-invasively track
and quantify the evolution of fluid flow over time (2-15 hours interval) and space (1-10 m
scale). The goal is to compare results from two field-scale experiments. The first
experiment was in the fracture-controlled Madonna della Mazza quarry. The second was
conducted in the undisturbed oolitic limestone in Ingraham Park. The purpose is to
assess the role of stratigraphic versus structural heterogeneities on fluid flow in highporosity carbonates. Furthermore, comparison between 4D GPR results and Eclipse
dynamic simulation offers insights to optimize workflows for more detailed flow models
and residual fluid recovery.
FIELD SITES DESCRIPTION
Madonna della Mazza quarry
The Madonna della Mazza quarry (MdM) is cut into the Orfento Formation situated on
the inner part of the Majella anticline (Southern Italy). The Upper Cretaceous formation
is composed of eroded subangular rudist fragments ranging in size from silt to rudite.
113
Porosity values range from 25% to 35% and permeability from 150 mD to 630 mD. The
stratigraphy of the quarry is characterized by prograding beds of grainstones
interbedded with thinner, fine-grained carbonate layers slightly dipping to the NE.
Previously conducted structural assessments of the entire quarry revealed the presence
of two main types of fractures: faults and deformation bands (Tondi et al., 2006).
Deformation bands are thin sheets of reduced porosity in the fault zone. These
cataclastic features form preferentially in the quarry uppermost, high-porosity, massive
grainstones due to micro-mechanical grinding of grains and do not show discontinuity
surfaces. Deformation bands generally have sealing properties for cross-fault fluid flow
and present porosities from 25% to 29%, lower than the grainstones with 32-35%. This
reduction of porosity results in a decrease of permeability up to half of the values
measured in the intact grainstones.
Ingraham Park
The field site of Ingraham Park is located southwest of Miami, between Coral Gables
and Coconut Grove. The park lies on a barrier oolitic bar which is part of a complex and
heterogeneous shoal system deposited during an interglacial Pleistocene sea-level highstand. Heterogeneities in the structurally undisturbed oolitic system are influenced by
grain size distribution, geometry of depositional bodies, and stratigraphy. The highly
variable rock matrix results in a wide variety of porosities ranging from 40% to 60% and
permeabilities from 600 mD to 1500 mD. Diagenesis and large-scale karstic dissolution
increase the degree of heterogeneity. Based on 3D GPR surveys covering the entire park
area, the 4D GPR site was selected to avoid the 1-4 m diameter filled-in dissolution holes
reported by Truss et al. (2007). Three main GPR facies, presenting high horizontal and
vertical variability, are observed in the 3D datasets: 1) bioturbated facies characterized
by worm burrows (lowermost); 2) perpendicular geometries interpreted as migrating
oolitic bars (middle); 3) cross-bedded continuous layers representing the prograding
shoal complex (uppermost). The majority of the connected porosity within the
uppermost reservoir unit is related to thin, coarse-grained shell hash beds typically
forming at the base of channels (Neal et al., 2008).
THE 4D GPR METHOD AND DATASETS
4D GPR is the acquisition of repeated 3D GPR surveys with identical geometries. The
purpose is to compare pairs of surveys and extract physical changes related to the fluid
migration while the surrounding matrix, not affected by water content changes, remains
unaltered.
In MdM 2952 liters of water were infiltrated from the surface into the host matrix over
a period of 30 hours. In Ingraham Park 3200 liters were infiltrated in 5 hours. Both
infiltrations were performed using 4 m diameter, temporary polyethylene pond walls
(circular for MdM, squared for Ingraham). The time intervals, measured after the end of
the infiltration, of repeated 3D GPR surveys used to compute snapshots of local water
content changes were: 1) 2-4 hours, 7-9 hours, 12-15 hours for MdM, and 2) 2-5 hours, 58 hours, 8-11 hours for Ingraham Park. Table 1 shows acquisition parameters and survey
information for the two field sites.
Table 1. Survey information for the MdM quarry and Ingraham Park field sites.
Field Site
GPR system, Frequency
MdM
Ingraham Park
Dual-channel, 200 MHz
Single-channel, 250 MHz
114
Inline
Spacing
5 cm
10 cm
Crossline
Spacing
2.5 cm
5 cm
Survey Volume
(X, Y, max. depth)
20 x 20 x ~12 m
18 x 20 x ~8 m
Water decreases the speed of electromagnetic waves and, as a consequence, increases
the traveltime of subsurface reflections. The application of a 3D crosscorrelation
algorithm (also known as WARP) allows automatic extraction of local timeshifts between
pairs of time-lapse surveys. Using the Topp petrophysical transfer function (Topp et al.,
1980) local water content changes can be calculated and displayed as a semi-transparent
attribute overlaying standard 3D GPR data (Fig. 1). Pairs of repeated surveys are used to
create snapshots of different stages during the waterbulb infiltration experiment. In each
snapshot the boundaries of the waterbulb are the draining zone (top) and the wetting
zone (bottom), the portions of matrix with negative and positive water content changes.
COMPARISON OF 4D GPR RESULTS FROM MDM AND INGRAHAM PARK
In MdM quarry structural features, such as deformation bands, play a key role in
influencing the fluid migration. Timeslices in Figure 1 (top part) show a pronounced
asymmetry of both draining (D) and wetting (W) zones, experiencing higher water
content changes over 2 hours in the undisturbed matrix (D: -3.5/-3%; W: +3.5/+4%)
than in the deformation bands area (D: -1.5/-1%; W: +1.5/+2%). Moreover, the wetting
zone shows a noticeable lateral distribution up-dip along the fault plane, indicating its
role as a preferential flow path (Marchesini et al., 2010).
Figure 1. Timeslices showing the evolution of draining (D) and wetting (W) zones for MdM (top part, 2-4
hours snapshot) and Ingraham Park (lower part, 2-5 hours snapshot). Draining zone is the upper
boundary of the waterbulb and wetting zone the lower. Depths of timeslices are marked in Figure 2. Black
circles (top part, for MdM) and black squares (lower part, for Ingraham Park) mark the boundaries of the
temporary, 4 m diameter polyethylene pond walls.
115
At Ingraham Park (Fig. 2, lower part) water content changes over 3 hours are larger
across the pond infiltration area in both draining and wetting zones (D: -5.5/-4.5%; W:
+6.5/+9%) suggesting a faster fluid migration compared to MdM. The upper waterbulb
boundary is within the pond perimeter while the lower boundary is shifted down-dip:
fluid migration in the wetting zone follows stratigraphy.
These observations are confirmed when comparing results of 4D GPR analysis in inline
direction near the center of the pond in Figure 2. Deformation bands in MdM (Fig. 2, top
part) compartmentalize the fluid while higher water content changes are experienced in
the undisturbed matrix. Stratigraphy has a lesser effect on fluid flow since both draining
(D) and wetting zone (W) do not dramatically exceed the vertical projection of the pond
into the subsurface.
On the contrary, water content changes in Ingraham Park are more laterally developed
and the shape of the water bulb follows stratigraphic boundaries (Fig. 2, lower part). As a
general trend also observed in timeslices, magnitudes of water content changes are
higher suggesting that higher porosity and permeability values of the host rock can be
related to more rapid fluid migration after the infiltration. The most lateral elongated
bodies of high water content changes follow shell hash beds and thin cemented calcite
layers (Grasmueck et al., 2007). The isolated water content change peak in the left
portion of the wetting zone is interpreted to be the result of out of plane fluid flow.
Figure 2. Inline sections corresponding to timeslices in Figure 1. Water bulb geometry in MdM (top part) is
affected by deformation bands while in Ingraham Park (lower part) draining and wetting zones mainly
follow stratigraphic layers.
116
HYDRAULIC HEADS AND WATER CONTENT CHANGES
Measurements of hydraulic heads over time, and in different locations within the pond
infiltration area, allow the fluid migration in the two different domains to be compared
and understood (Fig. 3). The hydraulic head of a moving water mass is a measure of the
gravitational force that causes groundwater to flow. A practical way to estimate this
quantity is to measure the height of the saturated water column or, in other words, the
local difference in depth between the bottom of the draining zone and the top of the
wetting zone. Hydraulic heads are measured at several XY locations within the pond
infiltration areas as the water content changes volumes. The results are plotted in Figure
3 against correspondent maximum volumetric water content changes measured in the
wetting zone of each hydraulic head evaluation point. Data show that in the case of MdM
there is a sharp difference in magnitude of water content changes between undisturbed
matrix and deformation bands, showing their role as fluid barriers for cross-fault fluid
migration, especially in fully-saturated conditions (i.e. the early stages of infiltration).
The influence of the deformation bands on fluid flow diminishes over time and water
content changes converge to similar magnitudes for lower values of hydraulic heads.
Figure 3. Measurements of hydraulic heads plotted against computed water content changes for MdM
(red to yellow) and Ingraham Park (shades of blue).
On the contrary, in Ingraham Park there are no separate trends of 4D GPR derived
water content changes and hydraulic head measurements, indicating homogeneous flow
across the entire pond area. Moreover, higher magnitudes of water content changes are
experienced when comparing all the three temporal snapshots: faster fluid migration is
achieved despite lower values of measured hydraulic heads. Less gravitational force is
needed to produce large water content changes suggesting that higher values of porosity
and permeability are responsible for faster fluid migration. This corroborates the
observations made for timeslices and inlines in Figure 1 and Figure 2. As a final remark,
hydraulic heads and correspondent water content changes plot on straight lines
indicating linear relationship between fluid velocity and pressure gradient, which is a key
condition to describe the infiltration mechanism as Darcy flow.
117
4D GPR RESULTS COMPARED WITH DYNAMIC FLUID FLOW MODELING
A static reservoir model of the surveyed portion of the MdM quarry was constructed
from detailed geological interpretation of the 3D GPR volumes and integration with
petrophysical parameters derived from sample plugs. This static model was the input for
a dynamic fluid flow simulation generated with Eclipse (Schlumberger Dynamic
Simulation Software) using the same infiltration settings as in the real MdM controlled
experiment presented in this study (2952 liters of water, 5 days infiltration time span).
As a preparation for the dynamic simulation, the original 3D stratigraphic and structural
interpretation had to be adapted in terms of: 1) precision: while strike orientation and
main geometries remained the same, small-scale heterogeneities (<30 cm) were lost.
Zig-zag shapes, adopted for both faults and deformation bands, did not preserve original
crosscutting relationships; 2) resolution: minimum cell size was set to 30 cm (total of
~150.000 cells) due to computational constrains while the original resolution of GPR
volumes is 5 cm (correspondent to ~4.000.000 cells). In addition, faults and
deformation bands have been set as zero-transmissibility surfaces in the orthogonal
direction. The simulation produced snapshots of absolute water saturation covering a
time span of 5 days with data reported every 12 hours (first snapshot: 2 hours after
infiltration).
Figure 4. Visualization of absolute values of water saturation from Eclipse dynamic fluid flow modeling
after 38 hours from the beginning of infiltration: inline section (top part) and timeslices (lower part) at
correspondent depths of draining and wetting zones as for the 4D GPR experiment.
118
There are differences when comparing results from 4D GPR method with dynamic fluid
modeling. The data show that in inline direction (Fig. 4, top part) the effect of
deformation bands on the fluid migration is completely lost: in the deformation bands
area higher values of water saturation are experienced while 4D GPR showed the
opposite behavior. The dynamic model fails to capture and visualize the role of structural
heterogeneities on the fluid behavior also in timeslice (Fig. 4, lower part) showing a
homogeneous distribution of water saturation across the whole pond area in both
sections corresponding to draining and wetting zones as described in Figure 1.
CONCLUSIONS AND IMPLICATIONS FOR RESERVOIR CHARACTERIZATION
The 4D GPR method was successfully applied to two reservoir analogues in gravity flow
experiments to visualize and quantify the effect of heterogeneities on fluid migration at
the 1-10 m scale and time intervals of 2-15 hours. The presented study shows different
fluid flow behaviors occurring in stratigraphic versus structural-controlled carbonate
domains offering insights to: 1) conduct more efficient reservoir characterization and 2)
to reduce uncertainties when upscaling from plug to field scale. Stratigraphic boundaries
have a relevant control on fluid migration in high-porosity, non-fractured domains. In
addition, structural heterogeneities (such as deformation bands) should be taken into
consideration when building static reservoir models for dynamic simulations. Realistic
flow models should always include both structural and stratigraphic heterogeneities in
order to improve reservoir kinematic studies and residual fluid recovery.
REFERENCES
Grasmueck, M., and D. Viggiano, 2007, Flowzone detection with time-lapse GPR water content
change measurements: CSL Annual Review Meeting.
Marchesini, P., Grasmueck, M., Eberli, G. P., and R. Van Dam, 2010, Tracking and quantifying
fluid flow in fractured Cretaceous carbonates with 4D Ground Penetrating Radar (GPR):
Madonna della Mazza Quarry, Italy: CSL Annual Review Meeting.
Neal, A., Grasmueck, M., McNeill, D. F., Viggiano, D. A., and G. P. Eberli, 2008, Full-resolution
3D radar stratigraphy of complex oolitic sedimentary architecture: Miami Limestone, Florida,
USA: Journal of Sedimentary Research, v. 78, p. 638-653.
Tondi, E., Antonellini, M., Aydin, A., Marchegiani, L., G. Cello, 2006, The role of deformation
bands, stylolites and sheared stylolites in fault development in carbonate grainstones of
Majella Mountain, Italy: Journal of Structural Geology, v. 28, p. 376-391.
Topp, G. C., Davis, J. L., and A. P. Annan, 1980. Electromagnetic determination of soil water
content: Measurements in coaxial transmission lines: Water Resource Research, v. 16, no. 3, p.
574-582.
Truss., S., Grasmueck, M., Vega, S., and D. A. Viggiano, 2007, Imaging rainfall drainage within
the Miami oolitic limestone using high-resolution time-lapse ground-penetrating radar: Water
Resource Research, v. 43, W0345, doi: 10.1029/2005WR004395.
119
120