ArticlePDF Available

Ocean Stagnation and end Permian anoxia

Authors:

Abstract

Ocean stagnation has been invoked to explain the widespread occurrence of organic-carbon-rich, laminated sediments interpreted to have been deposited under anoxic bottom waters at the time of the end-Permian mass extinction. However, to a first approximation, stagnation would severely reduce the upwelling supply of nutrients to the photic zone, reducing productivity. Moreover, it is not obvious that ocean stagnation can be achieved. Numerical experiments performed with a three-dimensional global ocean model linked to a biogeochemical model of phosphate and oxygen cycling indicate that a low equator to pole temperature gradient could have produced weak oceanic circulation and widespread anoxia in the Late Permian ocean. We find that polar warming and tropical cooling of sea-surface temperatures cause anoxia throughout the deep ocean as a result of both lower dissolved oxygen in bottom source waters and increased nutrient utilization. Buildup of quantities of H2S and CO2 in the Late Permian ocean sufficient to directly cause a mass extinction, however, would have required large increases in the oceanic nutrient inventory.
q2001 Geological Society of America. For permission to copy, contact Copyright Clearance Center at www.copyright.com or (978) 750-8400.
Geology; January 2001; v. 29; no. 1; p. 7–10; 3 figures; Data Repository item 20016. 7
Ocean stagnation and end-Permian anoxia
Roberta M. Hotinski* Department of Geosciences, Pennsylvania State University, University Park, Pennsylvania 16802, USA
Karen L. Bice Department of Geology and Geophysics, Woods Hole Oceanographic Institution, Woods Hole, Massachusetts
02543, USA
Lee R. Kump Department of Geosciences, Pennsylvania State University, University Park, Pennsylvania 16802, USA
Raymond G. Najjar Department of Meteorology, Pennsylvania State University, University Park, Pennsylvania 16802, USA
Michael A. Arthur Department of Geosciences, Pennsylvania State University, University Park, Pennsylvania 16802, USA
ABSTRACT
Ocean stagnation has been invoked to explain the widespread
occurrence of organic-carbon–rich, laminated sediments interpret-
ed to have been deposited under anoxic bottom waters at the time
of the end-Permian mass extinction. However, to a first approxi-
mation, stagnation would severely reduce the upwelling supply of
nutrients to the photic zone, reducing productivity. Moreover, it is
not obvious that ocean stagnation can be achieved. Numerical ex-
periments performed with a three-dimensional global ocean model
linked to a biogeochemical model of phosphate and oxygen cycling
indicate that a low equator to pole temperature gradient could
have produced weak oceanic circulation and widespread anoxia in
the Late Permian ocean. We find that polar warming and tropical
cooling of sea-surface temperatures cause anoxia throughout the
deep ocean as a result of both lower dissolved oxygen in bottom
source waters and increased nutrient utilization. Buildup of quan-
tities of H
2
S and CO
2
in the Late Permian ocean sufficient to di-
rectly cause a mass extinction, however, would have required large
increases in the oceanic nutrient inventory.
Keywords: Permian, anoxia, ocean circulation, phosphate cycle, oxy-
gen cycle.
INTRODUCTION
A number of authors have proposed that sluggish or stagnant
ocean circulation contributed to anoxia and the creation of large chem-
ical gradients in ancient oceans (e.g., Fischer and Arthur, 1977; Bra-
lower and Thierstein, 1984; Holser and Magaritz, 1987; Malkowski et
al., 1989; Gruszczynski et al., 1992; Kajiwara et al., 1994; Isozaki,
1997). Geochemical evidence suggests that the Permian-Triassic
boundary interval was such a period. Studies of boundary interval sed-
iments reveal large negative excursions in carbon, sulfur,and strontium
isotopic compositions of surface waters, which have been interpreted
as evidence of chemically distinct deep waters upwelling into a pre-
viously isolated surface ocean (Holser and Magaritz, 1987; Gruszcyn-
ski et al., 1992; Malkowski et al., 1994; Kajiwara et al., 1994; Knoll
et al., 1996). In addition, widespread laminated, pyritic sediments in
Upper Permian and Lower Triassic sections (Wignall and Hallam,
1992, 1993; Wignall and Twitchett, 1996; Isozaki, 1997) suggest that,
unlike today, decomposition of organic matter exhausted oxygen sup-
plied by circulation throughout much of the water column. The mag-
nitude and global nature of the excursions and anoxia led these inves-
tigators to propose that the entire ocean was severely stratified prior to
the boundary, and that stagnation-induced anoxia may have played a
role in the end-Permian extinction.
A potential flaw in this hypothesis is neglect of the relationship
between ocean circulation and surface productivity. While ocean mix-
ing acts to destroy vertical gradients in oxygen and other chemical
species, it also returns nutrients to the surface that fuel the biological
activity and organic-matter decomposition required to sustain gradients.
*Present address: Atmospheric and Ocean Sciences Program, Princeton
University, Princeton, New Jersey 08544, USA. E-mail: hotinski@princeton.
edu.
Quantifying the results of this competition between circulation and
biology thus requires a model that accounts for both ocean physics and
a dynamic link between nutrients and productivity. Although Permian
ocean circulation has been studied quantitatively in the past (Kutzbach
et al., 1990), this is the first attempt to treat the Permian ocean’s bio-
geochemistry explicitly.
We simulate a stagnation scenario for the Late Permian ocean
using a three-dimensional ocean general circulation model that includes
a simple biogeochemical model of phosphate and oxygen cycling. To
assess whether ocean stagnation would have created the chemical pat-
terns postulated for this time period, we apply a low equator-to-pole
temperature gradient and examine the changes in circulation and chem-
ical structure of the ocean that result.
MODEL DESCRIPTION
The general circulation model (GCM) used to simulate Permian
ocean circulation is the Geophysical Fluid Dynamics Laboratory’s
Modular Ocean Model (MOM) (Pacanowski et al., 1993). The model
configuration used includes 48348resolution, 16 vertical levels, and
constant vertical and horizontal mixing coefficients (1 cm
2
/s and 2 3
10
7
cm
2
/s, respectively). The ocean model bathymetry is a simplified
flat-bottom case (5150 m depth) with continental boundaries that ap-
proximate the land and sea distribution in the Late Permian (Wordian)
paleogeographical reconstruction of Rees et al. (1999). Further details
of the model, including additional simulations and model limitations,
can be obtained elsewhere.
1
For this study, the MOM code was modified to include simple
ocean biogeochemistry. Export of organic matter from the euphotic
zone (100 m thick) is modeled using the relationship of Yamanaka and
Tajika (1996). The export flux is proportional to surface-water phos-
phate concentration and varies with the cosine of latitude to simulate
light limitation with increasing latitude. Phosphate, rather than nitrate,
was chosen as the driver of productivity because phosphate is thought
to exert the primary control on marine primary production on long
time scales (e.g., Tyrrell, 1999). The exported organic flux is instan-
taneously remineralized below the euphotic zone according to the pow-
er law of Martin et al. (1987). Any flux that reaches the lowest vertical
layer of the ocean is remineralized, consistent with observations that
only a small fraction of the flux reaching the sediments is buried.
We assume that organic-matter decomposition will proceed via
denitrification and sulfate reduction after oxygen is depleted, and spec-
ify that the rate of phosphate remineralization is independent of oxygen
concentration in our model. However, because we do not include nitrate
and sulfate as tracers, such remineralization is represented by negative
values of dissolved O
2
. Although the biogeochemical model greatly
simplifies the cycling of organic matter, it produces realistic distribu-
tions of phosphate and oxygen in the modern ocean (Yamanaka and
Tajika, 1996).
To examine the effects of a reduced latitudinal temperature gra-
1
GSA Data Repository item 20016, Additional model description and pa-
rameters, is available on request from Documents Secretary, GSA, P.O. Box
9140, Boulder, CO 80301, editing@geosociety.org or at www.geosociety.org/
pubs/ft2001.htm.
8 GEOLOGY, January 2001
Figure 1. Model meridional (A) sea-surface temperature gradients and (B) surface density gradients for modern (A only), Permian high-
gradient, and Permian reduced-gradient scenarios.
dient on the Permian ocean, the model was run with two sets of cli-
matic forcings. A high-gradient forcing case used zonally averaged,
mean annual temperature, salinity, and wind forcings predicted by a
GENESIS version 2 simulation for the Late Permian (Wordian) that
included Late Permian paleogeography, 2760 ppm CO
2
(about eight
times the present atmospheric concentration), and 97.9% of the modern
solar luminosity (Rees et al., 1999). The temperature gradient predicted
by this simulation (Fig. 1A) is similar to the modern zonal average
gradient. The ocean model was run for 2700 yr with the high-gradient
forcing until a vigorous circulation was established.
To simulate climatic warming, the ocean model run wascontinued
with the same wind-stress and salinity forcing boundary conditions as
in the high-gradient case, but a new lower temperature gradient was
applied. Sea-surface temperatures (SSTs) for this simulation were fixed
at an equator-to-pole gradient of 12–28 8C (Fig. 1A), modeled after
that of the Paleocene-Eocene boundary interval (Zachos et al., 1994;
Bice et al., 2000). This gradient causes a substantial reduction in the
ocean’s pole to equator surface density gradient (Fig. 1B). Paleocli-
matological evidence (Taylor et al., 1992) suggests that the Late Perm-
ian southern polar climate was comparable to Paleocene-Eocene Arctic
climate, temperatures being .10 8C for at least a third of the year
(Ziegler, 1990; Yemane, 1993). In addition, oxygen isotope and paleo-
sol data indicate warming from the latest Permian into the earliest
Triassic (Holser et al., 1991; Retallack, 1999).
RESULTS
Circulation
The steady-state meridional overturning (zonally integrated mass
transport stream function in the meridional vertical plane) for the high-
gradient scenario is shown in Figure 2A. The pattern is asymmetrical;
there is a strong cell in the Southern Hemisphere and a weaker cell in
the Northern Hemisphere. The maximum Southern Hemisphere trans-
port value is .80 Sverdrups (1 Sv 51310
6
m
3
s
2
1
) and the circu-
lation can be characterized as vigorous.
This vigorous circulation is disrupted when the reduced temper-
ature gradient is applied (Fig. 2). At 100 yr, the ocean is poorly mixed
and is dominated below 1000 m by Northern Hemisphere sinking. By
1200 yr, weak mixing between the surface and deep ocean is reestab-
lished. After 10 k.y., circulation has reached a new steady state with
substantially reduced overturning relative to the high-gradient case,
consistent with the reduction in upper ocean density between the sim-
ulations. The steady state thermohaline circulation, driven by a weaker,
but largely symmetrical, density contrast (Fig. 1B), exhibits much less
interhemispheric asymmetry than the high-gradient circulation (Fig.
1A). Although circulation is reduced, bottom water is still formed at
high latitudes in both hemispheres and there is no low-latitude deep-
water formation. Northern Hemisphere deep waters are ventilated at a
rate nearly comparable to that of modern North Atlantic Deep Water,
estimated to be ;20 Sv (Broecker, 1991).
Dissolved Oxygen
The high-gradient case is characterized by high values of dis-
solved O
2
in the high latitudes and in deep water (Fig. 3). The ocean
is oxic everywhere except in intermediate waters off the western coast
of Pangea (Fig. 3A), where negative values of O
2
indicate oxidation
of organic matter with electron acceptors other than oxygen (i.e., nitrate
and sulfate). The deep ocean is particularly well oxygenated (Fig. 3B);
minimum O
2
values are .150 mmol/L. In comparison, modern North
Pacific deep waters average ;130 mmol/L.
When the ocean reaches a steady state after application of the
reduced temperature gradient, oxygen concentrations below the top 100
m are reduced by an average of 264 mmol/L. Deep ocean oxygen levels
are dysoxic (,45 mmol/L) to anoxic over a majority of the deep-sea
floor (Fig. 3D), and intermediate water oxygen concentrations are neg-
ative; values are as low as 2300 mmol/L off the western coast of the
supercontinent (Fig. 3C).
Knoll et al. (1996) discussed a box model of Permian anoxia and
suggested that utilization of ;470 mmol/L of sulfate in the Late Perm-
ian caused buildup of lethal CO
2
levels in the deep ocean and contrib-
uted to the end-Permian extinction. Because sulfate has twice the re-
ducing power of oxygen, this amount of sulfate reduction would
correspond to values of ;21000 mmol/L oxygen in the model. Such
levels are not even approached in the reduced-gradient simulation.
Therefore, some factor in addition to reduced overturning seems nec-
essary to create such high levels of CO
2
in the Late Permian deep
ocean.
DISCUSSION AND CONCLUSIONS
Our results suggest that deep-ocean anoxia is consistent with re-
duced thermohaline circulation driven by a low meridional density con-
trast. However, almost half the difference between the simulated high-
gradient and reduced-gradient deep water oxygen levels is due to the
difference in solubility of oxygen in the warmed high-latitude regions
where deep waters are formed (;250 mmol/L, vs. 370 mmol/L in the
GEOLOGY, January 2001 9
Figure 2. Simulated me-
ridional overturning (zon-
ally integrated volume
transport stream func-
tion in meridional vertical
plane) for Permian ocean
in Sverdrups (Sv; 10
6
m
3
/
s). Positive values indi-
cate clockwise flow, and
negative values indicate
counterclockwise flow.
Modular ocean model
(MOM) (A) equilibrated
with high-gradient sea-
surface temperature forc-
ing, and (B) 100 yr, (C)
1200 yr, and (D) 10 k.y. af-
ter imposing reduced-
gradient forcing.
Figure 3. Maps of steady-
state oxygen concentrations
(mmol/L) for (A) intermediate
(850 m) and (B) deep ocean
(4650 m) in high-gradient
scenario; (C) upper interme-
diate and (D) deep ocean in
reduced-gradient scenario.
10 GEOLOGY, January 2001
high-gradient scenario). This result is consistent with box modeling
results for the Cretaceous (Herbert and Sarmiento, 1991). Biological
oxygen demand explains the remainder of the oxygen decline (;140
mmol/L), but the magnitude of this oxygen demand alone would not
drive the reduced-gradient ocean completely anoxic without the afore-
mentioned temperature effect. Chemical stratification is limited because
decreased upwelling rates of nutrients in the reduced-gradient case sup-
port lower export of organic matter from surface waters than in the
high-gradient case (14.3 vs. 19.8 Gt C/yr). The effects of slowed ven-
tilation and dwindling productivity do not exactly balance because the
surface biota exploit a reserve source of nutrients, i.e., high-latitude
phosphate.
Because of light limitation at high latitudes, the high-gradient sce-
nario with vigorous upwelling exhibits phosphate concentrations .2
mmol/L in southern high latitudes and 1.5 mmol/L in the northern areas
of convection. In the reduced-gradient scenario, with slower upwelling
and longer high-latitude surface-water residence times, phosphate val-
ues in surface waters of both the northern and southern high latitudes
decline by ;0.4 and 0.8 mmol/L, respectively (not shown). As a result,
3.62 310
20
mol of phosphate are lost from surface waters in the
reduced-gradient case, fueling export of organic matter that drives al-
ready low deep-water oxygen concentrations to anoxic levels. This sen-
sitivity of deep-water oxygenation to high-latitude productivity is con-
sistent with the box-modeling results of Sarmiento et al. (1988) and
Hotinski et al. (2000). Thus, although the reduced-gradient scenario
exhibits widespread anoxia, relatively high rates of ventilation and re-
duced productivity keep the deep ocean near the oxic-anoxic boundary,
rather than firmly in the regime of sulfate reduction. Increasing the
marine phosphate inventory by 50% increases chemical stratification
in the model (results not shown), but five times more phosphate would
be needed to drive the system to the values suggested by Knoll et al.
(1996). Because the phosphate content of the Permian ocean is not
well determined, such a high value cannot be ruled out (Martin, 1995).
High steady-state phosphate concentrations in pore waters of modern
marine sediments, which commonly approach 10 mmol/L, indicate that
high levels of phosphate are sustainable (Emerson et al., 1980; Van
Cappellen and Berner, 1988). Thus, it is possible that the Permian
ocean’s phosphate concentration was significantly higher than today’s
value of 2.15 mmol/L.
ACKNOWLEDGMENTS
Financial support was provided by the NASA Astrobiology Institute Cooperative
Agreement (NCC2-1057), the National Science Foundation (grant EAR-98-05139), the
Shell Oil Company Foundation, and the Pennsylvania Space Grant Consortium. We thank
Tim Bralower for a thoughtful review of the manuscript.
REFERENCES CITED
Bice, K.L., Sloan, L.C., and Barron, E.J., 2000, Comparison of early Eocene isotopic
paleotemperatures and the three-dimensional OGCM temperature field: The potential
for use of model-derived surface water d
18
O, in Huber, B.T., et al., eds., Warm
climates in Earth history: Cambridge, UK, Cambridge University Press, p. 79–131.
Bralower, T.J., and Thierstein, H.R., 1984, Low productivity and slow deep-water circu-
lation in mid-Cretaceous oceans: Geology, v. 12, p. 614–618.
Broecker, W.S., 1991, The great ocean conveyor: Oceanography, v. 4, p. 79–89.
Emerson, S., Jahnke, R., Bender, M., Froelich, P., Klinkhammer, G., Bowser, C., and Set-
lock, S., 1980, Early diagenesis in sediments from the eastern Equatorial Pacific, I.
Pore water nutrient and carbonate results: Earth and Planetary Science Letters, v. 49,
p. 57–80.
Fischer, A.G., and Arthur, M.A., 1977, Secular variations in the pelagic realm, in Cook,
H.E., and Enos, P., eds., Deep-water carbonate environments: Society of Economic
Paleontologists and Mineralogists Special Publication 25, p. 19–50.
Gruszczynski, M., Hoffman, A., Malkowski, K., and Veizer, J., 1992, Seawater strontium
perturbation at the Permian-Triassic boundary, West Spitsbergen, and its implications
for the interpretation of strontium isotopic data: Geology, v. 20, p. 779–782.
Herbert, G.D., and Sarmiento, J.L., 1991, Ocean nutrient distribution and oxygenation:
Limits on the formation of warm saline bottom waters over the past 91 m.y.: Ge-
ology, v. 19, p. 702–705.
Holser, W.T., and Magaritz, M., 1987, Events near the Permian-Triassic boundary: Modern
Geology, v. 11, p. 155–180.
Holser, W.T., Scho¨nlaub, H.P., Boeckelman, K., and Magaritz, M., 1991, The Permian-
Triassic of the Gartnerkofel-1 core (Carnic Alps, Austria):Synthesis and conclusions,
in Holser, W.T., and Scho¨nlaub, H.P., eds., The Permian-Triassic boundary in the
Carnic Alps of Austria (Gartnerkofel region): Abhandlungen Geologische Bunde-
santstalt Austriche, v. 45, p. 213–232.
Hotinski, R.M., Kump, L.R., and Najjar, R.G., 2000, Opening Pandora’s Box: The impact
of open-system modeling on interpretations of anoxia: Paleoceanography, v. 15,
p. 267–279.
Isozaki, Y., 1997, Permo-Triassic boundary superanoxia and stratified superocean:Records
from lost deep-sea: Science, v. 276, p. 235–238.
Kajiwara, Y., Yamakita, S., Ishida, K., Ishiga, H., and Imai, A., 1994, Development of a
largely anoxic stratified ocean and its temporary massive mixing at the Permian/
Triassic boundary supported by the sulfur isotopic record: Palaeogeography,Palaeo-
climatology, Palaeoecology, v. 111, p. 367–379.
Knoll, A.H., Bambach, R.K., Canfield, D.E., and Grotzinger, J.P., 1996, Comparative Earth
history and Late Permian mass extinction: Science, v. 273, p. 452–457.
Kutzbach, J.E., Guetter, P.J., and Washington, W.M., 1990, Simulated circulation of an
idealized ocean for Pangaean time: Paleoceanography, v. 5, p. 299–317.
Malkowski, K., Gruszczynski, M., Hoffman, A., and Halas, S., 1989, Oceanic stable iso-
topic composition and a scenario for the Permo-Triassic crisis: Historical Biology,
v. 2, p. 289–309.
Martin, J.H., Knauer, G.A., Karl, D.M., and Broenkow, W.W., 1987, VERTEX: Carbon
cycling in the northeast Pacific: Deep Sea Research, v. 32, p. 267–286.
Martin, R.E., 1995, Cyclic and secular variation in microfossil mineralization, clues to the
biogeochemical evolution of Phanerozoic oceans: Global and Planetary Change,
v. 11, p. 1–23.
Pacanowski, R., Dixon, K., and Rosati, A., 1993, The Geophysical Fluid Dynamics Lab-
oratory modular ocean users guide: Geophysical Fluid Dynamics Laboratory Ocean
Group Technical Report 2.
Rees, P.M., Gibbs, M.T., Ziegler, A.M., Kutzbach, J.E., and Behling, P.J., 1999, Permian
climates; evaluating model predictions using global paleobotanical data: Geology,
v. 27, p. 891–894.
Retallack, G., 1999, Postapocalyptic greenhouse paleoclimate revealed by earliest Triassic
paleosols in the Sydney Basin, Australia: Geological Society of America Bulletin,
v. 111, p. 52–70.
Sarmiento, J.L., Herbert, T.D., and Toggweiler, J.R., 1988, Causes of anoxia in the world
ocean: Global Biogeochemical Cycles, v. 7, p. 417–450.
Taylor, E.L., Taylor, T.N., and Cu´ neo, N.R., 1992, The present is not the key to the past:
A polar forest from the Permian of Antarctica: Science, v. 257, p. 1675–1677.
Tyrrell, T., 1999, The relative influences of nitrogen and phosphorus on oceanic primary
production: Nature, v. 400, p. 525–531.
Van Cappellen, P., and Berner, R.A., 1988, A mathematical model for the early diagenesis
of phosphorus and fluorine in marine sediments: Apatite precipitation: American
Journal of Science, v. 288, p. 289–333.
Wignall, P.B., and Hallam, A., 1992, Anoxia as a cause of the Permian/Triassic mass
extinction, facies evidence from northern Italy and the western United States: Pa-
laeogeography, Palaeoclimatology, Palaeoecology, v. 93, p. 21–46.
Wignall, P.B., and Hallam, A., 1993, Griesbachian (Early Triassic) palaeoenvironmental
changes in the Salt Range, Pakistan and southeast China and their bearing on the
Permo-Triassic extinction: Palaeogeogragraphy, Palaeoclimatology, Palaeoecology,
v. 102, p. 215–237.
Wignall, P.B., and Twitchett, R.J., 1996, Oceanic anoxia and the end Permian mass ex-
tinction: Science, v. 272, p. 1155–1158.
Yamanaka, Y., and Tajika, E., 1996, The role of the vertical fluxes of particulate organic
matter and calcite in the oceanic carbon cycle; studies using an ocean biogeochemical
general circulation model: Global Biogeochemical Cycles, v. 10, p. 361–382.
Yemane, K., 1993, Contribution of Late Permian palaeogeography in maintaining a tem-
perate climate in Gondwana: Nature, v. 361, p. 51–54.
Zachos, J.C., Stott, L., and Lohmann, K.C., 1994, Evolution of early Cenozoic marine
temperatures: Paleoceanography, v. 9, p. 353–387.
Ziegler, A.M., 1990, Phytogeographic patterns and continental configurations during the
Permian Period, in McKerrow, W.S., and Scotese, C.R., eds., Palaeozoic palaeo-
geography and biogeography and climatology: Geological Society [London]Memoir
12, p. 363–379.
Manuscript received May 1, 2000
Revised manuscript received September 21, 2000
Manuscript accepted October 4, 2000
Printed in USA
... Euxinic is an additional term used to describe anoxic water masses with hydrogen sulphide (Allison et al., 1995;Lyons & Severmann, 2006). Ocean anoxia is thought to be one of the driving killing mechanisms in the Early Triassic, and has been shown to be an overarching control on recovery patterns after the end-Permian extinction (EPME) (e.g., Wignall & Twitchett, 1996;Hotinski et al., 2001;Bottjer et al., 2008;Clapham & Payne, 2011;Pietsch & Bottjer, 2014;Song et al., 2014;Bond & Grasby, 2017;Benton, 2018;Zhang et al., 2018). ...
... Because the solubility of gases decreases with increasing temperature, SST exerts a first-order control on dissolved oxygen concentrations in seawater. Moreover, warming likely resulted in a sluggish meridional overturning circulation, reducing upwelling, mixing and ventilation (Hotinski et al., 2001;Fig. 8. Ce-anomalies, Yb/Nd, Y/Ho and corrected U contents of the carbonate fraction with litho-, bio-and chemostratigraphic framework for the Chanakhchi section. ...
Article
Ocean anoxia was one of the key killing mechanisms responsible for the end-Permian mass extinction (∼252 Ma). However, the temporal evolution and the triggering mechanisms of the end-Permian anoxia are controversial, with the current view being that the water column deoxygenation was a spatially and temporally heterogeneous event. Here, we use cerium-anomalies, uranium contents and rare earth element and yttrium (REY) compositions measured on the carbonate fraction of samples from two marine sections in Armenia and South China to constrain the evolution of end-Permian marine anoxia. In particular, we attempt to identify intervals of manganous redox conditions (i.e., Mn²⁺-rich water column). These are characterized by the reduction of Mn(IV)-oxides at a redox potential lower than of nitrate and higher than of Fe(III)-oxide reduction. Most of our samples yield REY patterns with seawater characteristics suggesting the dominance of hydrogenous REYs. While decreasing U concentrations in the C. yini Zone suggest an overall intensification of global marine anoxia in the latest Permian, Ce-anomalies shift from negative to positive above the C. yini Zone, implying a deterioration in the oxygen content of the local water columns. Highly positive Ce-anomalies (Ce/Ce* > 1.3) are firstly reported for the P-T interval, indicating enrichment of Ce in the water column and the establishment of manganous redox conditions at both locations. This change coincided globally with climatic warming, the onset of the marine extinction interval as well as locally with the disappearance of sponge spicules and radiolarians (South China) and the appearance of microbialites (Armenia).
... Under these conditions, an upward shift of the redoxcline would have expanded the volume of the ocean in which anaerobic denitrification and anammox reactions could occur, while simultaneously reducing the oceanic volume in which nitrification could take place, resulting in a net consumption of seawater NO 3 − . Oceanic stagnation and stratification would have further inhibited deep-sea N recycling, disrupting the balance between source and sink fluxes of nutrient N (Hotinski et al., 2001). As a consequence, fixation of atmospheric N 2 became the dominant source of bioavailable nitrogen in the ocean surface layer (Fig. 6A-B). ...
Article
Global warming, widespread oceanic anoxia and stagnation, and large perturbations of the global carbon cycle characterized the end-Permian to Middle Triassic interval. Nitrogen isotopes of marine sediments (δ15Nbulk) decreased through the Permian–Triassic transition, implying development of nitrate-limited and ammonium-dominated conditions (i.e., anaerobic marine N cycle) in Early Triassic oceans, which may have contributed to the delay in marine biotic recovery following the end-Permian mass extinction. However, the temporal evolution of the nitrogen cycle and the role of nutrient supply in marine ecosystem recovery during this interval remain poorly understand. Here, we present a new high-resolution Permian–Triassic nitrogen isotope curve from the Nanpanjiang Basin of South China. Low δ15N (−2‰ to +2‰) during the Griesbachian-to-Smithian substages (i.e., first ∼2 Myr of Early Triassic) reflects enhanced nitrogen fixation concurrently with climatic hyperwarming and expansion of oceanic anoxia. A large rise in δ15N (to +8‰) followed by a decline (to −2‰) reflects an aborted recovery of the marine N cycle during the Spathian substage of the Early Triassic. During the Middle Triassic, δ15N fluctuations between +1‰ and +4‰ during the Anisian Stage, followed by stabilization around +4‰ in the Ladinian Stage, suggest a slow stepwise re-establishment of the aerobic marine N cycle. Although both South China and northwestern Pangea experienced a transition to anaerobic N cycling during the Early Triassic, South China experienced an earlier and more rapid onset of this event as well as larger N-cycle fluctuations during the recovery interval than northwestern Pangea. Overall, N cycle changes coincided with those in paleotemperature and ocean-redox state, demonstrating an integrated response of the marine system to the extreme environmental perturbations of the Early Triassic. In summary, our results support persistence of nutrient-N-limited conditions and strong microbial N-fixation throughout the Early Triassic, with a full return to the aerobic N cycle only after stabilization of oceanic environmental conditions during the Middle Triassic.
... Ocean stagnation, a decrease in the overturning circulation, is often invoked to explain the occurrence of widespread black shales and anoxic conditions in Earth's history (Bralower and Thierstein, 1984;Gruszczynski et al., 1992). However, results from more sophisticated models demonstrate that ocean stagnation may not induce anoxia, where the supply of nutrients to the surface ocean is insufficient to support the elevated productivity required to sustain high O 2 demand in the deep ocean (Hotinski et al., 2001). Another perspective is that eustatic changes trigger variation in nutrient levels and primary productivity, which can lead to changes in organic carbon fluxes and deepwater redox conditions, independent of the extent of deepwater restriction (Arthur and Sageman, 2005;Liu et al., 2014). ...
Article
The early Middle Triassic (Anisian) Luoping biota is representative of marine ecosystems following their full recovery after the Permian-Triassic mass extinction; however, its exceptional preservation remains poorly understood. In this paper, we report multiple geochemical proxies (TOC, TN, P/Al, Cu xs , Ni xs , Ba xs , δ 13 C carb , and δ 13 C org) from Member II of the Middle Triassic Guanling Formation at the Xiangdongpo section in Luoping County, Yunnan Province, China, to assess the relationship between primary productivity and exceptional preservation of the Luoping biota. Variations in TOC, TN, P/Al, Cu xs , and Ni xs values show that exceptional preservation coincides with two intervals of high productivity. Relatively low C/N ratios and distributed Δ 13 C carb-org values indicate that primary productivity was dominated by eukaryotic algae and prokaryotic microbes. Based on geochemical analyses and regional correlations, we conclude that, following sea level rise, nutrients were supplied to the Luoping area by upwelling from the open ocean and facilitated the blooming of marine communities in an open platform setting. In the intra-platform depression, increased nutrients supply also fueled high primary productivity in surface waters and oxygen consumption in the water column, causing bottom water anoxia. The lack of oxygen in the bottom water reduced the rate of biomass degradation and bioturbation, promoting growth of microbial mats, leading to the exceptional preservation of macrofauna. This study highlights the important role of elevated primary productivity in exceptional fossil preservation by triggering anoxia of bottom waters. Our findings confirm the widespread development of anoxic conditions during the middle Anisian (Pelsonian), from the eastern Tethys to the Panthalassa, and also reveal the interactions between organisms and the environment after Permian-Triassic mass extinction.
... With regard to nutrient availability, we limited our review to published data that showed a more or less protracted decrease of terrestrial sediment input during the course of the Paleozoic (Flögel et al., 2000;Hay et al., 2006;Prokoph et al., 2008), and a decline of upwelling of nutrients in the late Paleozoic (Parrish, 1982(Parrish, , 1987Hotinski et al., 2001). By comparing these data with our diversity curves of the marine phytoplankton, we observe a possible positive response to terrestrial derived nutrient input: High diversity occurred in intervals of high nutrient availability, such as during the Ordovician, when acritarchs had their highest diversity. ...
Article
In our recent publication on the biodiversity of Paleozoic phytoplankton (Kroeck et al., 2022) we cited several works of Martin (Martin, 1996; Martin and Quigg, 2012; Martin and Servais, 2020) to present the hypothesis that during the late Paleozoic nutrient availability in the oceans increased significantly from a rather (super-) oligotrophic early to middle Paleozoic. We put this view in contrast to the notion based on data on terrestrial sediment input that nutrient availability generally decreased over the course of the Paleozoic. In his discussion paper, Martin (2022) expands on this hypothesis and the role of nutrients for biodiversification, and clarifies the use of the term “superoligotrophy” (Martin, 1996) for the early to middle Paleozoic interval. Martin (2022) highlights factors we neglected in our review of marine phytoplankton diversity in the Paleozoic. We largely agree with the arguments presented, and therefore consider this discussion as an excellent addition to our review.
... While organic-rich phosphatic black shales during Anisian times are still characterized by a 15 N content typical of newly fixed atmospheric N 2 (∼0‰) due to replenished surface nutrients with low N:P ratio 5 (Fig. 8), the progressive rise of N-isotopes from −1‰ to +3‰ during the Ladinian (∼242-237 Ma) (Fig. 8) suggests increasing upwelling of a partially denitrified water mass into the photic zone and subsequent utilization of various plankton groups. It marked the end of ocean stagnation with high nutrient inventories, as simulated for the LPE 85 , however, not yet tested for the Early to Middle Triassic. Our data of gradual recovery from ocean nutrient stress since the latest Anisian corroborate a period of atmospheric CO 2 drawdown, reduced radiative forcing, and associated ocean cooling (Fig. 10) and reflect improved ocean mixing. ...
Article
Full-text available
Following the latest Permian extinction ∼252 million years ago, normal marine and terrestrial ecosystems did not recover for another 5-9 million years. The driver(s) for the Early Triassic biotic crisis, marked by high atmospheric CO2 concentration, extreme ocean warming, and marine anoxia, remains unclear. Here we constrain the timing of authigenic K-bearing mineral formation extracted from supergene weathering profiles of NW-Pangea by Argon geochronology, to demonstrate that an accelerated hydrological cycle causing intense chemical alteration of the continents occurred between ∼254 and 248 Ma, and continued throughout the Triassic period. We show that enhanced ocean nutrient supply from this intense continental weathering did not trigger increased ocean productivity during the Early Triassic biotic crisis, due to strong thermal ocean stratification off NW Pangea. Nitrogen isotope constraints suggest, instead, that full recovery from ocean nutrient stress, despite some brief amelioration ∼1.5 million years after the latest Permian extinction, did not commence until climate cooling revitalized the global upwelling systems and ocean mixing ∼10 million years after the mass extinction.
... However, it is difficult to assess the significance of these upwelling zones for phytoplankton evolution, but they might have been an important source for nutrients, and hotspots for primary production, in the early Paleozoic. A decrease of phosphorite deposition suggests a decline of upwelling of nutrients in the late Paleozoic (Parrish, 1982(Parrish, , 1987Hotinski et al., 2001), coinciding with the apparent low phytoplankton diversity during this interval. ...
Article
Phytoplankton form the base of most marine trophic chains and studying their past diversity at regional and global scales can provide valuable insights into the evolution of marine ecosystems and climate history. Using a new database of more than 4000 species of acritarchs and prasinophytes, a comprehensive investigation of the taxonomic diversity trajectories of this marine (micro)phytoplankton throughout the Paleozoic is performed for the first time. This dataset compiles data from published literature, including taxonomic, geographic and stratigraphic information at the stage resolution. Our results highlight five major temporal trends in phytoplankton diversity variation: (i) an initial plateau of moderate richness during the early and middle Cambrian, followed by (ii) a sharp increase from the late Cambrian to the Middle Ordovician, which records the highest Paleozoic diversity of organic-walled phytoplankton (OWP); then, (iii) a protracted decrease during the Late Ordovician to Middle Devonian; (iv) a slight peak in diversity during the Late Devonian, before (v) falling to the lowest richness recorded during the Carboniferous and Permian. The role of phytoplankton during major biotic events is discussed: While phytoplankton evolution may have been a factor in enabling the “Cambrian Explosion”, we do not find a strong relationship between the diversity changes of the phytoplankton and this event and we thus refute the notion that it might have been a major driver of radiations during this interval. However, a strong increase in phytoplankton diversity coincides with the Great Ordovician Biodiversification Event (GOBE), indicating that the profound changes of marine phytoplankton, and thus of the base of marine food webs, enabled diversifications throughout marine ecosystems. A decrease in phytoplankton diversity during the Lower and Middle Devonian points against the hypothesis of phytoplankton triggering the proposed “Devonian Nekton Revolution”. By comparing the results with paleoenvironmental parameters, several factors are found to be possibly related to the long-term diversity trends: Our results indicate that paleogeography and sea-level changes were probably the main drivers of phytoplankton diversity patterns throughout the Paleozoic, while increases in sediment influx provided facilitating conditions for phytoplankton diversification. Atmospheric CO2 concentration as well as temperature and related sea ice cover are found to be further important controlling factors for phytoplankton diversity.
... As a result, tropical temperatures surged from ~25 • C prior to the eruption to nearly 40 • C. This "Super Hothouse" global warming parched life on land and led to the formation of an anoxic "Strangelove" ocean (Hsu et al., 1985;Kump, 1991;Hotinski et al., 2001;Zhang et al., 2001;Grice et al., 2005;Heydari et al., 2008). ...
Article
This study provides a comprehensive and quantitative estimate of how global temperatures have changed during the last 540 million years. It combines paleotemperature measurements determined from oxygen isotopes with broader insights obtained from the changing distribution of lithologic indicators of climate, such as coals, evaporites, calcretes, reefs, and bauxite deposits. The waxing and waning of the Earth’s great polar icecaps have been mapped using the past distribution of tillites, dropstones, and glendonites. The global temperature model presented here includes estimates of global average temperate (GAT), changing tropical temperatures (ΔT◦ tropical), deep ocean temperatures, and polar temperatures. Though similar, in many respects, to the temperature history deduced directly from the study of oxygen isotopes, our model does not predict the extreme high temperatures for the Early Paleozoic required by isotopic investigations. The history of global changes in temperature during the Phanerozoic has been summarized in a “paleotemperature timescale” that subdivides the many past climatic events into 8 major climate modes; each climate mode is made up of 3-4 pairs of warming and cooling episodes (chronotemps). A detailed narrative describes how these past temperature events have been affected by geological processes such as the eruption of Large Igneous Provinces (LIPS) (warming) and bolide impacts (cooling). The paleotemperature model presented here allows for a deeper understanding of the interconnected geologic, tectonic, paleoclimatic, paleoceanographic, and evolutionary events that have shaped our planet, and we make explicit predictions about the Earth’s past temperature that can be tested and evaluated. By quantitatively describing the pattern of paleotemperature change through time, we may be able to gain important insights into the history of the Earth System and the fundamental causes of climate change on geological timescales. These insights can help us better understand the problems and challenges that we face as a result of Future Global Warming.
Article
Full-text available
Brachiopod shell samples from the Kapp Starostin Formation, West Spitsbergen, provide evidence for a large and rapid drop in seawater strontium isotopic ratios close to the Permian-Triassic boundary. The strontium isotopic shift is associated with similarly dramatic declines in carbon and oxygen isotope curves recorded in the same samples. This pattern can be explained by a paleoceanographic model in which there is replacement of a largely stagnant, stratified ocean by a vigorously mixed one at the Permian-Triassic boundary. -Authors
Article
Full-text available
A seven-component upper ocean ecosystem model of nitrogen cycling calibrated with observations at Bermuda Station “S” has been coupled to a three-dimensional seasonal general circulation model (GCM) of the North Atlantic ocean. The aim of this project is to improve our understanding of the role of upper ocean biological processes in controlling surface chemical distributions, and to develop approaches for assimilating large data sets relevant to this problem. A comparison of model predicted chlorophyll with satellite coastal zone color scanner observations shows that the ecosystem model is capable of responding realistically to a variety of physical forcing environments. Most of the discrepancies identified are due to problems with the GCM model. The new production predicted by the model is equivalent to 2 to 2.8 mol m−2 yr−1 of carbon uptake, or 8 to 12 GtC/yr on a global scale. The southern half of the subtropical gyre is the only major region of the model with almost complete surface nitrate removal (nitrate
Article
Full-text available
The most recent global icehouse-hothouse climate transition in Earth history occurred in the Permian. Warmer polar conditions relative to today existed from the middle Permian through the Mesozoic and into the Cenozoic. We focus here on one particularly well-correlated middle Permian stage that postdated the deglaciation, the Wordian (267-264 Ma), integrating floral and lithological data to determine Wordian climates globally. Paleobotanical data provide the best means of interpreting terrestrial paleoclimates, often revealing important information in the con- tinuum between "dry" and "wet" end-member lithological indicators such as evaporites and coals. New statistical analyses of Wordian floras worldwide have enabled a greater understand- ing of original vegetation patterns and prevailing climate conditions. The derived climate inter- pretations are compared with new Wordian atmospheric general circulation model simulations. The model matches the data well in the tropics and northern high latitudes, but predicts colder conditions in southern high latitudes. We discuss possible reasons for this discrepancy.
Article
The processes included in the modeling are decomposition of organic matter, dissolution of biogenic mineral debris, reduction of hydrous ferric oxides, authigenic carbonate fluorapatite (CARFAP) precipitation, reversible adsorption/desorption of dissolved P, molecular diffusion, and sediment burial. Applications of the model to the different types of P and F diagenesis are discussed. -Authors
Article
Distributions of chemical tracers in the world ocean are well reproduced in an ocean general circulation model which includes biogeochemical processes (biogeochemical general circulation model, B-GCM). The difference in concentration of tracers between the surface and the deep water depends not only on the export production but also on the remineralization depth. Case studies changing the vertical profile of particulate organic matter (POM) flux and the export production show that the phosphate distribution can be reproduced only when the vertical profile of POM flux observed by sediment traps is used. The export production consistent with the observed distribution of phosphate is estimated to be about 10 GtC/yr. Case studies changing the vertical profile of calcite flux and the rain ratio, a ratio of production rate of calcite against that of particulate organic carbon (POC), show that the rain ratio should be smaller than the widely used value of 0.25. The rain ratio consistent with the observed distribution of alkalinity is estimated to be 0.08 to approximately 0.10. This value can be easily understood in a two-box model where the difference of remineralization depth between POC and calcite is taken into account.
Article
Facies, faunal and geochemical evidence from the Permian/Triassic boundary sediments of the Dolomites and Idaho indicates a major anoxic event in the earliest Triassic. In both regions, the basal beds consist of finely laminated micrites with common syngenetic pyrite. The only fauna consists of occasional bedding plane assemblages of Lingula or Clacaria, a typical lower dysaerobic assemblage. This is a level where previous studies have shown a major negative carbon isotope excursion and a cerium anomaly. In the Dolomites, the pyritic micrite directly overlies strata containing a diverse and typically Permian marine fauna of algae, foraminifers (including fusulinids) and articulate brachiopods, implying an abrupt extinction in contradiction to many previous views.Sequence stratigraphic analysis of the Dolomite boundary sediments reveals a minor sequence boundary in the late Permian followed by extremely rapid transgression leading to the development of the relatively deep water pyritic micrite — a maximum flooding surface at the Permo-Triassic boundary. A further pulsed deepening in the lower Griesbachian, recorded in both the Dolomites and Idaho, lead to the widespread establishment of dysaerobic facies. It is clear that most of the extinctions occurred at the erathem boundary although the subsequent failure of the marine benthos to fill the empty ecospace in the ensuing Griesbachian may have been due to the widespread development of dysaerobic conditions.
Article
The geologic record preserves evidence that vast regions of ancient oceans were once anoxic, with oxygen levels too low to sustain animal life. Because anoxic conditions have been postulated to foster deposition of petroleum source rocks and have been implicated as a kill mechanism in extinction events, the genesis of such anoxia has been an area of intense study. Most previous models of ocean oxygen cycling proposed, however, have either been qualitative or used closed-system approaches. We reexamine the question of anoxia in open-system box models in order to test the applicability of closed-system results over long timescales and find that open and closed-system modeling results may differ significantly on both short and long timescales. We also compare a scenario with basinwide diffuse upwelling (a three-box model) to a model with upwelling concentrated in the Southern Ocean (a four-box model). While a three-box modeling approach shows that only changes in high-latitude convective mixing rate and character of deepwater sources are likely to cause anoxia, four-box model experiments indicate that slowing of thermohaline circulation, a reduction in wind-driven upwelling, and changes in high-latitude export production may also cause dysoxia or anoxia in part of the deep ocean on long timescales. These results suggest that box models must capture the open-system and vertically stratified nature of the ocean to allow meaningful interpretations of long-lived episodes of anoxia.