Vous êtes sur la page 1sur 9

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/303380996

Carbonate Factories

Chapter · January 2016


DOI: 10.1007/978-94-007-6238-1_136

CITATION READS

1 1,484

1 author:

John J. G. Reijmer
King Fahd University of Petroleum and Minerals, Dhahran, Saudi Arabia
196 PUBLICATIONS   2,684 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Cold water corals View project

Oligo-Miocene carbonates from the Indo-Pacific View project

All content following this page was uploaded by John J. G. Reijmer on 25 May 2016.

The user has requested enhancement of the downloaded file.


Encyclopedia of Marine Geosciences
DOI 10.1007/978-94-007-6644-0_136-1
# Springer Science+Business Media Dordrecht 2014

Carbonate Factories
John J.G. Reijmer*
Department of Sedimentology and Marine Geology, Faculty of Earth and Life Sciences (FALW), VU University
Amsterdam, Amsterdam, The Netherlands

Synonyms
Carbonate production systems

Definition
Carbonate factories, or production systems, are benthic carbonate associations that show variations
in their dominant precipitation mode, mineral composition, and depth range of production as well as
growth potential (Schlager, 2000).

Introduction
The term “carbonate factory” was introduced to define the narrow depth zone where tropical reefs
and detrital carbonates are produced (e.g., Tucker and Wright, 1990; James and Kendall, 1992;
Reading and Levell, 1996). Based on the carbonate factory principle, Schlager (2000) proposed a
threefold subdivision of the benthic carbonate production systems, with the planktonic carbonate
factory as a fourth system. The latter is traditionally dealt with in the context of paleo-oceanography.
The Schlager (2000) carbonate factory concept is based on the style of carbonate precipitation in
aquatic realms: (1) abiotic, (2) biotically induced, or (3) biotically controlled (Lowenstam, 1981). In
the latter category, a distinction can be made between sunlight-controlled organisms (phototrophic)
and nutrient-controlled organisms (heterotrophic).

Environmental Parameters
A series of environmental parameters steer the different modes of carbonate precipitation. The
abiotic mode is normally encountered in marine and freshwater aquatic settings, although increasing
evidence suggests microbes also play a mediating role in the formation of whitings (e.g., Yates and
Robbins, 1998; Thompson, 2000) and ooids (Pacton et al., 2012). The differentiation found in the
biotically controlled precipitation mode is directed through a series of environmental factors
that drive environmental variations, the most important factors being: (1) light, (2) temperature,
(3) nutrients, and (4) salinity. These factors set the boundaries for styles of carbonate precipitation
and sediment production profiles but also for sediment production and distribution. As a result, the

*Email: j.j.g.reijmer@vu.nl
*Email: john.reijmer@falw.vu.nl

Page 1 of 8
Encyclopedia of Marine Geosciences
DOI 10.1007/978-94-007-6644-0_136-1
# Springer Science+Business Media Dordrecht 2014

overall morphological development of the carbonate system relates to the dominance of specific
factors:

1. Light. This is considered as one of the most important environmental controls. The depth of the
photic column, light penetration, varies around the present-day globe with a maximum of almost
150 m for the Pacific atolls (Schlager, 2005).
The light-saturated zone and euphotic zone not only regulate the growth forms of corals but
most importantly the growth rates of the photosynthetic, carbonate-secreting benthos (Schlager,
2003).
2. Temperature. More or less equal to light is temperature, as it regulates the diversity of the biotic
association. Each carbonate-secreting species has its own optimum growth window along the
temperature scale (Lees, 1975).
Ocean circulation patterns and latitudinal positions relate to the distribution and occurrence of
specific water temperatures and hence determine the distribution of specific organisms across the
globe.
3. Nutrients. Nutrient variations play an important role as high nutrient levels reduce calcification
rates, but also stimulate the development of filamentous algae, bryozoans, and barnacles (e.g.,
Halfar et al., 2004, 2006; Reijmer et al., 2012); increase bio-erosion rates (Chazottes et al., 2008);
and may hamper coral recruitment (Smith and Buddemeier, 1992; Atkinson et al., 1995).
4. Salinity. Coral communities are tolerant to long-term and short-term salinity variations (Muthiga
and Szmant, 1987). The same holds for Mytilus shells (Malone and Dodd, 1967) whose
calcification rates are not influenced by salinity thresholds. Algal communities exist that have a
greater tolerance for high salinities, and they occur in large terminal lakes like the Great Salt Lake
(Stephens, 1990; Harris et al., 2013). Porter et al. (1999) showed, for the Florida Keys coral reef
system, that elevated salinities could diminish the negative effects of elevated temperature and
conclude that temperature and salinity have opposing effects on coral photosynthesis. In
restricted environments, however, biotic diversity might be reduced (Miocene, SE Spain; e.g.,
Braga and Martin, 1996).

The influence of the aforementioned factors on the growth strategies of skeletal and microbial
precipitation and on the formation of non-skeletal particles differs for each component. Various
strategies will result in different products at different times. Schlager (2000, 2003, 2005) distin-
guished three end-members of carbonate sedimentation systems or the so-called carbonate factories.
The specific factories possess different sedimentation modes that reflect different growth strategies
with different overall geometries, carbonate mineralogy, and grain sizes.

Carbonate Factories
The three main factories (Fig. 1) that were distinguished are:
(1) T-factory, in which the T is derived from tropical or “top-of-the-water-column” (Schlager,
2005); (2) the C-factory, in which the C stands for cool-water or controlled precipitation; and (3) the
M-factory, in which M represents microbial, micrite, or mud-mound (Schlager, 2003, 2005).
A fourth factory might be distinguished, the cold-water reef systems, which share characteristics
with the T-factory through the type of dominant skeletal builder, e.g., scleractinian corals, and with
the C-factory type, because of the production-depth profile and the nutrient-steered and light-
independent carbonate production mode.

Page 2 of 8
Encyclopedia of Marine Geosciences
DOI 10.1007/978-94-007-6644-0_136-1
# Springer Science+Business Media Dordrecht 2014

MARINE
CARBONATE
PRECIPITATION FACTORIES

abiotic biotic

biotically induced biotically controlled

heterotrophic autotrophic

MUD-MOUND COOL-WATER TROPICAL


COLD-WATER
CORALS

Fig. 1 Classification scheme of carbonate factories related to their precipitation modes (see text for details) (Modified
from Schlager (2003))

Depositional geometries and sediment export modes Production rates and production depth

TROPICAL TROPICAL COOL-WATER MUD-MOUND


Shoreline 0 1 0 1 0 1
0m

100
Spread
Genuine
drowning
possible.
COOL-WATER 200

300
Move

MUD-MOUND
400

Localized
production on Production
500
current-swept around
deep-sea methane seeps
floors. in deep sea.
Stick

Fig. 2 Platform morphologies, production rates, and sediment export mode of the individual carbonate factories
(Modified from Schlager (2003))

Page 3 of 8
Encyclopedia of Marine Geosciences
DOI 10.1007/978-94-007-6644-0_136-1
# Springer Science+Business Media Dordrecht 2014

In the T-factory, light and water temperature steer the production profiles ensuring high produc-
tion rates through biota living in the photic zone (Fig. 2). The occurrence of this factory is mostly
limited to the tropical zone between 30 N and 30 S, with modification through surface currents
related to ocean gyres and upwelling areas.
The light and water temperature steered production mode results in a carbonate platform mor-
phology, with a rim, reef barrier, at the edge of the platform (Fig. 2).This barrier protects the shallow-
water lagoon environments and forms the upper part of the steep slopes surrounding the platform.
These slopes are mostly coarse grained and show fast cementation (Grammer et al., 1999).
The major mineralogy here is aragonite explaining the fast cementation rates found on the slopes
(Grammer et al., 1999), reef systems, platform interior, and exposed sediments (Dravis, 1996).
This factory is very sensitive to relative changes in sea level as it is closely tied to the light-
saturated zone, and shows a flat top and steep slopes. A relatively small drop in sea level might
eliminate most of the sediment production zone. This sedimentation pattern with sea-level-related
sharp changes in sediment production is called highstand shedding (e.g., Schlager et al., 1994). One
could tag the overall sedimentation pattern as “Spread” as the systems export abundant sediment
from shallow-water areas into surrounding basins (Fig. 2).
Nutrients and relatively low temperatures steer the biologically controlled production of carbon-
ate sediments in the C-factory. In surface waters, the C-factory operates at higher latitudes than the
T-factory, northward of 30 N and southward of 30 S. Other realms are upwelling areas and
low-temperature waters below the thermocline.
The production profile shows moderate production rates over a large water depth range (Fig. 2), so
production is not limited to shallow waters (Schlager, 2003). Another characteristic is the open
sedimentary system without any shallow-water barriers as found in the other factories. Low
cementation rates combined with the absence of a shallow-water barrier result in a sedimentation
regime in which waves and currents control sediment transport and redistribution. One could label
the overall sedimentation pattern as “Move” as changes in currents and waves result in a shift of the
depositional sediment loci. In addition, the sedimentary response to sea-level changes shows large
similarities to that seen in clastic systems (Fig. 2).
The M-factory is characterized by the microbial-mediated, e.g., bacteria and cyanobacteria,
precipitation of mud (Schlager, 2000, 2003). The most prominent features of this factory are the
microbial mats, which usually contain several microbial species from various key groups, and their
organic EPS (extracellular polymeric substances) matrix steering different organo-mineralization
processes. The production profile of this factory shows a fairly steady production rate (Fig. 2), which
can reach tropical rates (Kenter et al., 2005), within a wide range of environments and water depths.
Very characteristic for this factory is the formation of mound buildups, e.g., the Carboniferous of
Belgium (Lees and Miller, 1995) and Devonian of Algeria (Wendt et al., 1997), albeit large flat-
topped platforms are also very frequent, e.g., the Carboniferous of Kazakhstan (Kenter et al., 2005)
and Triassic of the Dolomites (Keim and Schlager, 2001). The production profile shows high
production rates down to 500 m water depth or even deeper, e.g., around methane seeps (Fig. 2).
Production in the shallow-water areas is affected by the waves due to the sensitivity of the biota to
these physical processes.
The term “Stick” describes the overall sedimentation pattern very well, because of the ability of
the system to glue (through microbial activity) sediments and build sturdy systems (Fig. 2).
Relative changes in sea level do not have a large impact on the production rates of this system as
the main sediment production loci are situated on the upper slope, below the wave base, and extend
over a fairly extensive water depth range. As a result, the production rates of this system are fairly

Page 4 of 8
Encyclopedia of Marine Geosciences
DOI 10.1007/978-94-007-6644-0_136-1
# Springer Science+Business Media Dordrecht 2014

constant and can be described by slope shedding (Kenter et al., 2005), slope-derived sediment
production not being affected by sea-level changes.
As a fourth system, one could add the cold-water coral reefs (CWCR), first described from the
Northwestern European margin by Teichert (1958). These systems share the stony corals with the
T-factory and the nutrient-dependent growth strategy with the C-factory. Characteristic for this
CWCR-factory is the dominance of ahermatypic corals, e.g., Lophelia and Madrepora, in a
sedimentary system with very high biodiversity. The nutrient dependency and occurrence
throughout a wide range of water depths below the photic zone result in the worldwide distribution
of this sedimentary system, with new discoveries being added daily (Correa et al., 2012;
Mienis et al., 2012).

Production Profiles
The production profiles of the different factories not only reflect the water depths in which the
factories reach their optimum growth and sediment production, but also the environmental processes
that steer the types and amounts of sediment produced. Various environmental factors drive the
development of specific carbonate producing organisms and associated modes of carbonate pro-
duction, which results in differences in the grain-size spectra produced, as well as dominant
carbonate mineralogies. These factors then relate to the efficiency of the system to build sturdy
structures, fixing the sediments within depositional realms.
So the different styles of carbonate platform development are related to production profiles and
thus the sensitivity of the system as a whole to environmental changes. These differences are very
well expressed in the response of the T-, C-, and M-factories to relative changes in sea level. Sharp
contrasts are found in the T-factory, with high production during highstands in sea level when the flat
tops of the platforms are flooded. Sharply reduced production occurs during times when the tops
become exposed and sediment production is restricted to small surfaces on the steep platform slopes.
In the C-factory, relative changes in sea level have a minor effect on the overall production rates, but
do cause re-sedimentation of sediments during these changes. The M-factory also shows fairly stable
sediment production and export rates during sea-level changes, as the main sediment production
sites are situated on the upper slope. Minor differences might occur during phases of progradation
and aggradation (Della Porta et al., 2004). The CWCR-factory strongly depends on the steady influx
of nutrients either through slope currents or pelagic input. The CWCR-factories along the Irish
margin as well as the communities in the Mediterranean show variations in their occurrence related
to glacial and interglacial time intervals.

Conclusion/Summary
The carbonate factory concept provides a subdivision of marine aquatic benthic carbonate sediment
production systems based on their styles of carbonate precipitation. The planktonic carbonate
factory is classified as an additional production system. The overall carbonate sediment production

Page 5 of 8
Encyclopedia of Marine Geosciences
DOI 10.1007/978-94-007-6644-0_136-1
# Springer Science+Business Media Dordrecht 2014

for the individual systems (tropical, cool-water, microbial, and cold-water coral reef systems)
depends on specific factory-dependent sediment production profiles.

Cross-References
▶ Chemosynthetic Life
▶ Eustasy
▶ Export Production
▶ Foraminifers (Benthic)
▶ Guyot, Atoll
▶ Lagoon
▶ Lithostratigraphy
▶ Marine Sedimentary Basins
▶ Reef Coast
▶ Reefs (Rock and Coral)
▶ Sea-Level Dynamics
▶ Sediment Transport Models
▶ Sedimentary Sequence/Parasequence
▶ Sequence Stratigraphy
▶ Systems Tracts (Highstand, Lowstand, Transgressive, Shelf Margin)

Bibliography
Atkinson, M. J., Carlson, B., and Crow, G. L., 1995. Coral growth in high-nutrient, low-ph seawater:
a case study of corals cultured at Waikiki Aquarium, Honolulu, Hawaii. Coral Reefs, 14(4),
215–223.
Braga, J. C., and Martín, J. M., 1996. Geometries of reef advance in response to relative sea-level
changes in a Messinian (uppermost Miocene) fringing reef (Cariatiz reef, Sorbas Basin, SE
Spain). Sedimentary Geology, 107, 61–81.
Chazottes, V., Reijmer, J. J. G., and Cordier, E., 2008. Sediment characteristics in reef areas
influenced by eutrophication-related alterations of benthic communities and bioerosion processes.
Marine Geology, 250(1–2), 114–127.
Correa, T. B. S., Grasmueck, M., Eberli, G. P., Reed, J. K., Verwer, K., and Purkis, S., 2012.
Variability of cold-water coral mounds in a high sediment input and tidal current regime, Straits of
Florida. Sedimentology, 59(4), 1278–1304.
Della Porta, G., Kenter, J. A. M., and Bahamonde, J. R., 2004. Depositional facies and stratal
geometry of an Upper Carboniferous prograding and aggrading high-relief carbonate platform
(Cantabrian Mountains, N. Spain). Sedimentology, 51, 267–295.
Dravis, J. J., 1996. Rapidity of freshwater calcite cementation – implications for carbonate
diagenesis and sequence stratigraphy. Sedimentary Geology, 107, 1–10.
Grammer, G. M., Crescini, C. M., McNeill, D. F., and Taylor, L. H., 1999. Quantifying rates of
syndepositional marine cementation in deeper platform environments – new insight into a
fundamental process. Journal of Sedimentary Research, 69(1), 202–207.

Page 6 of 8
Encyclopedia of Marine Geosciences
DOI 10.1007/978-94-007-6644-0_136-1
# Springer Science+Business Media Dordrecht 2014

Halfar, J., Godinez-Orta, L., Mutti, M., Valdez-Holguín, J. E., and Borges, J. M., 2004. Nutrient and
temperature controls on modern carbonate production. An example from the Gulf of California,
Mexico. Geology, 32(3), 213–216.
Halfar, J., Godinez-Orta, L., Mutti, M., Valdez-Holguin, J. E., and Borges, J. M., 2006. Carbonates
calibrated against oceanographic parameters along a latitudinal transect in the Gulf of California,
Mexico. Sedimentology, 53, 297–320.
Harris, P. M., Ellis, J., and Purkis, S. J., 2013. Assessing the extent of carbonate deposition in early
rift settings. American Association of Petroleum Geologists Bulletin, 97(1), 27–60.
James, N. P., and Kendall, A. C., 1992. Introduction to carbonate and evaporite facies models. In
Walker, R. G., and James, N. P. (eds.), Facies Models – Response to Sea Level Change. St. John’s:
Geological Society of Canada, pp. 265–275.
Keim, L., and Schlager, W., 2001. Quantitative compositional analysis of a Triassic carbonate
platform (Southern Alps, Italy). Sedimentary Geology, 139, 261–283.
Kenter, J. A. M., Harris, P. M., and Della Porta, G., 2005. Steep microbial boundstone-dominated
platform margins – examples and implications. Sedimentary Geology, 178, 5–30.
Lees, A., 1975. Possible influence of salinity and temperatures on modern shelf carbonate sedimen-
tation. Marine Geology, 19, 159–198.
Lees, A., and Miller, J., 1995. Waulsortian banks. In Monty, C. L. V., Bosence, D. W. J., Bridges,
P. H., Pratt, B. R. (eds.), Carbonate Mud-Mounds – Their Origin and Evolution. International
Association of Sedimentologists, Oxford-London (UK), pp. 191–271.
Lowenstam, H. A., 1981. Minerals formed by organisms. Science, 211, 1126–1131.
Malone, P. G., and Dodd, J. R., 1967. Temperature and salinity effects on calcification rate in
Mytilus edulis and its paleoecological implicatons. Limnology and Oceanography, 12(3),
432–436.
Mienis, F., Duineveld, G. C. A., Davies, A. J., Ross, S. W., Seim, H., Bane, J., and Van Weering,
T. C. E., 2012. The influence of near-bed hydrodynamic conditions on cold-water corals in the
Viosca Knoll area, Gulf of Mexico. Deep Sea Research, Part I, 60, 32–45.
Muthiga, N. A., and Szmant, A. M., 1987. The effect of salinity stress on the rates of aerobic
respiration and photosynthesis in the hermatypic coral Siderastrea Siderea. The Biological
Bulletin, 173, 539–551.
Pacton, M., Ariztegui, D., Wacey, D., Kilburn, M. R., Rollion-Bard, C., Farah, R., and Vasconcelos,
C., 2012. Going nano: a new step toward understanding the processes governing freshwater ooid
formation. Geology, 40, 547–550.
Porter, J. W., Lewis, S. K., and Porter, K. G., 1999. The effect of multiple stressors on the Florida
keys coral reef ecosystem: a landscape hypothesis and a physiological test. Limnology and
Oceanography, 44(3), 941–949.
Reading, H. G., and Levell, B. K., 1996. Controls on the sedimentary rock record. In Reading, H. G.
(ed.), Sedimentary Environments: Processes, Facies and Stratigraphy, 3rd edn. Oxford, UK:
Blackwell, pp. 5–52.
Reijmer, J. J. G., Bauch, T., and Sch€afer, P., 2012. Carbonate facies patterns in surface sediments of
upwelling and non-upwelling shelf environments (Panama, East Pacific). Sedimentology, 59(1),
32–56.

Page 7 of 8
Encyclopedia of Marine Geosciences
DOI 10.1007/978-94-007-6644-0_136-1
# Springer Science+Business Media Dordrecht 2014

Schlager, W., 2000. Sedimentation rates and growth potential of tropical, cool-water and
mud-mound carbonate systems. In Insalaco, E., Skelton, P. W., and Palmer, T. J. (eds.), Carbonate
Platform Systems: Components and Interactions. London: The Geological Society, pp. 217–227.
Schlager, W., 2003. Benthic carbonate factories of the Phanerozoic. International Journal of Earth
Sciences, 92, 445–464.
Schlager, W., 2005. Carbonate Sedimentology and Sequence Stratigraphy. Tulsa: SEPM (Society
for Sedimentary Geology). SEPM Concepts in Sedimentology and Paleontology, Vol. 8. 200 pp.
Schlager, W., Reijmer, J. J. G., and Droxler, A. W., 1994. Highstand shedding of carbonate
platforms. Journal of Sedimentary Research, B64(3), 270–281.
Smith, S. V., and Buddemeier, R. W., 1992. Global change and coral reef ecosystems. Annual
Review of Ecology and Systematics, 23, 89–118.
Stephens, D. W., 1990. Changes in lake levels, salinity and the biological community of Great Salt
Lake (Utah, USA), 1847–1987. Hydrobiologia, 197, 139–146.
Teichert, C., 1958. Cold- and deep-water coral banks. American Association of Petroleum Geolo-
gists Bulletin, 42(5), 1064–1082.
Thompson, J. B., 2000. Microbial whitings. In Riding, R. E., and Awramik, S. M. (eds.), Microbial
Sediments. Berlin/Heidelberg: Springer, pp. 250–260.
Tucker, M. E., and Wright, V. P., 1990. Carbonate Sedimentology. Oxford, UK: Blackwell. 482 pp.
Wendt, J., Belka, Z., Kaufmann, B., Kostrewa, R., and Hayer, J., 1997. The world’s most spectacular
carbonate mud mounds (Middle Devonian, Algerian Sahara). Journal of Sedimentary Research,
67(3), 424–436.
Yates, K. K., and Robbins, L. L., 1998. Production of carbonate sediments by a unicellular green
alga. American Mineralogist, 83, 1503–1509.

Page 8 of 8

View publication stats

Vous aimerez peut-être aussi