extinguishing a permian world - wordpress.com · are fusulinid foraminifera, rugose and tabulate...

2
287 At the end of the Permian, ca. 252 Ma ago, marine and terrestrial fauna were facing the most extensive mass extinction in Earth history (Raup and Sepkoski, 1982). 80%–95% of all species on Earth, on land and in the oceans, became extinct (Benton et al., 2004) within an estimated time interval of less than 200 k.y. to 700 k.y. (Huang et al., 2011; Shen et al., 2011). Among the prominent Paleozoic animal groups that vanished are fusulinid foraminifera, rugose and tabulate corals, and the arthropod class Trilobita. The numerous hypotheses about the causes of the mass extinction include various environmental changes, mostly related to the emplacement of the Siberian Traps large igneous province. The compila- tion of radiometric U/Pb ages for the mass extinction and the Siberian Traps demonstrate a temporal overlap of both events (Svensen et al., 2009). A prominent hypothesis for the mass extinction is an accentuated global climate change scenario induced by volcanic CO 2 degassing (e.g., Svensen et al., 2009) that triggered biotic responses in the sea and on land. But what do we know about the climate at this time in Earth history? The two main features of climate recorded in the geological archives are temperature and humidity i.e., the moisture that is available for plant growth, and to a certain extent rainfall patterns. Climate simulations model the general circulation patterns during Permian–Triassic times. The paleogeography was characterized by the supercontinent Pangea extend- ing nearly from pole to pole, with large land masses in the mid-latitudes of the Northern and Southern Hemispheres, and with the Tethys Ocean in the tropics (e.g., Smith et al., 1994). Climate models and sensitivity experi- ments demonstrated that this paleogeography provided the preconditions for a monsoonal circulation with strong seasonality of temperatures and rainfall on the Tethyan coasts, and distinct northern and southern intertropi- cal convergence zones (Fig. 1) (e.g., Kutzbach and Ziegler, 1993; Parrish, 1993). Moist conditions prevailed in middle and high latitudes and along the western Pangea coast, contrasting with the year-round arid Pangean tropics. (Kutzbach and Ziegler, 1993; Parrish, 1993). Recent climate modeling for the Permian–Triassic scenario tested the impact of rapid temperature change and showed that shifts in biomes were less pronounced during global warm- ing compared to global cooling (Roscher et al., 2011). For the Late Perm- ian oceans, sensitivity experiments with coupled atmosphere-ocean models imply that elevated CO 2 levels caused significant temperature increases in high-latitude sea-surface water masses, leading to reduced global sea-sur- face temperature (SST) gradients (e.g., Kiehl and Shields, 2005). A proxy record for ancient SSTs is stable oxygen isotope data of marine biomineralizing fauna. During biomineralization, marine fauna incorporate dissolved oxygen into their hard parts (e.g., calcite, aragonite, apatite) in or close to isotopic equilibrium with the ambient sea water (Urey, 1947; Epstein et al., 1951). Therefore, geochemical signals preserved in marine invertebrate shells and skeletal parts can be used as a proxy for SST, if the isotope composition of the ambient water is known (although isotope frac- tionation during biomineralization and diagenetic alteration need to be con- sidered). Oxygen isotopes measured on unaltered brachiopod shells from the Late Permian Bellerophon Formation, and from a single sample from the overlaying basal Werfen Formation, in southern Italy were interpreted to represent a rise of 6–10 °C of tropical SST (Kearsey et al., 2009 and references therein). However, the resolution and stratigraphic constraints of these data are limited. In the search for unaltered archives for oxygen iso- tope records for paleotemperature reconstructions, the significance of bio- genic apatite of tooth enamel is increasing. The oxygen isotope signature of conodont apatite is little affected by post-depositional changes (Joachimski and Buggisch, 2002). New analytical techniques and better understanding of isotopic changes during diagenesis have revealed the importance of the conodont paleothermometry for the reconstruction of ancient ocean tem- peratures (e.g., Barham et al., 2012; Vennemann et al., 2002). The work of Joachimski et al. (2012, p. 195 in this issue of Geology) uses the conodont paleothermometry to reconstruct ocean temperatures of a tropical site during the largest mass extinction in Earth’s history. One of the best studied Permian–Triassic successions is the Global Boundary Stratotype Section and Point (GSSP) in Meishan, South China. Joachim- ski et al. measured the oxygen isotope values of P1 elements of the con- odont genera Clarkina (or Neogondolella) and Hindeodus recovered from the Meishan and Shangsi sections, choosing a δ 18 O value of –1‰ Vienna standard mean ocean water for their temperature calculation. They demon- strate that in the Meishan area, SST increased by 1–5 °C, from an average temperature of 22 °C, before the main extinction event. Thereafter, tem- peratures remained fairly stable up to the extinction event, where oxygen isotopes indicate another distinct warming by 5–8 °C, up to SST of 32–35 °C in the earliest Triassic. The data of Joachimski et al. are an important advance in the discussion of the Permian–Triassic climate change, as they quantify SST change in a stratigraphically well-constrained framework. The documented oxygen isotope shift (i.e., the warming) coincides with a negative shift in carbon isotope ratios over a period of possibly only ~110 k.y. (Shen et al., 2011). Geology, March 2012; v. 40; no. 3; p. 287–288; doi: 10.1130/focus032012.1. © 2012 Geological Society of America. For permission to copy, contact Copyright Permissions, GSA, or [email protected]. Figure 1. Permian–Triassic Pangean paleogeography, modified af- ter Smith et al. (1994). Paleogeographic position of South China and Meishan (star). Precipitation/evaporation ratio (P < E, P > E) after Ziegler et al. (2003). Late Permian biomes and northern and southern intertropical convergence zones (NITC, SITC) modified after Kutzbach and Ziegler (1993). Extinguishing a Permian World Elke Schneebeli-Hermann Palaeoecology, Department of Physical Geography, Utrecht University, Budapestlaan 4, 3584 CD Utrecht, Netherlands NITC SITC E>P E>P E<P E<P Land Mountain Warm temperate humid Tropical humid Subtropical arid Tropical semi humid Cold temperate Cool temperate South China Tethys Ocean

Upload: others

Post on 01-Feb-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Extinguishing a Permian World - WordPress.com · are fusulinid foraminifera, rugose and tabulate corals, and the arthropod class Trilobita. The numerous hypotheses about the causes

GEOLOGY, March 2012 287

At the end of the Permian, ca. 252 Ma ago, marine and terrestrial fauna were facing the most extensive mass extinction in Earth history (Raup and Sepkoski, 1982). 80%–95% of all species on Earth, on land and in the oceans, became extinct (Benton et al., 2004) within an estimated time interval of less than 200 k.y. to 700 k.y. (Huang et al., 2011; Shen et al., 2011). Among the prominent Paleozoic animal groups that vanished are fusulinid foraminifera, rugose and tabulate corals, and the arthropod class Trilobita. The numerous hypotheses about the causes of the mass extinction include various environmental changes, mostly related to the emplacement of the Siberian Traps large igneous province. The compila-tion of radiometric U/Pb ages for the mass extinction and the Siberian Traps demonstrate a temporal overlap of both events (Svensen et al., 2009). A prominent hypothesis for the mass extinction is an accentuated global climate change scenario induced by volcanic CO2 degassing (e.g., Svensen et al., 2009) that triggered biotic responses in the sea and on land. But what do we know about the climate at this time in Earth history?

The two main features of climate recorded in the geological archives are temperature and humidity i.e., the moisture that is available for plant growth, and to a certain extent rainfall patterns. Climate simulations model the general circulation patterns during Permian–Triassic times. The paleogeography was characterized by the supercontinent Pangea extend-ing nearly from pole to pole, with large land masses in the mid-latitudes of the Northern and Southern Hemispheres, and with the Tethys Ocean in the tropics (e.g., Smith et al., 1994). Climate models and sensitivity experi-ments demonstrated that this paleogeography provided the preconditions for a monsoonal circulation with strong seasonality of temperatures and rainfall on the Tethyan coasts, and distinct northern and southern intertropi-cal convergence zones (Fig. 1) (e.g., Kutzbach and Ziegler, 1993; Parrish, 1993). Moist conditions prevailed in middle and high latitudes and along the western Pangea coast, contrasting with the year-round arid Pangean tropics. (Kutzbach and Ziegler, 1993; Parrish, 1993). Recent climate modeling for the Permian–Triassic scenario tested the impact of rapid temperature change and showed that shifts in biomes were less pronounced during global warm-ing compared to global cooling (Roscher et al., 2011). For the Late Perm-ian oceans, sensitivity experiments with coupled atmosphere-ocean models imply that elevated CO2 levels caused signifi cant temperature increases in high-latitude sea-surface water masses, leading to reduced global sea-sur-face temperature (SST) gradients (e.g., Kiehl and Shields, 2005).

A proxy record for ancient SSTs is stable oxygen isotope data of marine biomineralizing fauna. During biomineralization, marine fauna incorporate dissolved oxygen into their hard parts (e.g., calcite, aragonite, apatite) in or close to isotopic equilibrium with the ambient sea water (Urey, 1947; Epstein et al., 1951). Therefore, geochemical signals preserved in marine invertebrate shells and skeletal parts can be used as a proxy for SST, if the isotope composition of the ambient water is known (although isotope frac-tionation during biomineralization and diagenetic alteration need to be con-sidered). Oxygen isotopes measured on unaltered brachiopod shells from the Late Permian Bellerophon Formation, and from a single sample from the overlaying basal Werfen Formation, in southern Italy were interpreted to represent a rise of 6–10 °C of tropical SST (Kearsey et al., 2009 and references therein). However, the resolution and stratigraphic constraints of these data are limited. In the search for unaltered archives for oxygen iso-tope records for paleotemperature reconstructions, the signifi cance of bio-

genic apatite of tooth enamel is increasing. The oxygen isotope signature of conodont apatite is little affected by post-depositional changes (Joachimski and Buggisch, 2002). New analytical techniques and better understanding of isotopic changes during diagenesis have revealed the importance of the conodont paleothermometry for the reconstruction of ancient ocean tem-peratures (e.g., Barham et al., 2012; Vennemann et al., 2002).

The work of Joachimski et al. (2012, p. 195 in this issue of Geology) uses the conodont paleothermometry to reconstruct ocean temperatures of a tropical site during the largest mass extinction in Earth’s history. One of the best studied Permian–Triassic successions is the Global Boundary Stratotype Section and Point (GSSP) in Meishan, South China. Joachim-ski et al. measured the oxygen isotope values of P1 elements of the con-odont genera Clarkina (or Neogondolella) and Hindeodus recovered from the Meishan and Shangsi sections, choosing a δ18O value of –1‰ Vienna standard mean ocean water for their temperature calculation. They demon-strate that in the Meishan area, SST increased by 1–5 °C, from an average temperature of 22 °C, before the main extinction event. Thereafter, tem-peratures remained fairly stable up to the extinction event, where oxygen isotopes indicate another distinct warming by 5–8 °C, up to SST of 32–35 °C in the earliest Triassic. The data of Joachimski et al. are an important advance in the discussion of the Permian–Triassic climate change, as they quantify SST change in a stratigraphically well-constrained framework. The documented oxygen isotope shift (i.e., the warming) coincides with a negative shift in carbon isotope ratios over a period of possibly only ~110 k.y. (Shen et al., 2011).

Geology, March 2012; v. 40; no. 3; p. 287–288; doi: 10.1130/focus032012.1.© 2012 Geological Society of America. For permission to copy, contact Copyright Permissions, GSA, or [email protected].

Figure 1. Permian–Triassic Pangean paleogeography, modifi ed af-ter Smith et al. (1994). Paleogeographic position of South China and Meishan (star). Precipitation/evaporation ratio (P < E, P > E) after Ziegler et al. (2003). Late Permian biomes and northern and southern intertropical convergence zones (NITC, SITC) modifi ed after Kutzbach and Ziegler (1993).

Extinguishing a Permian WorldElke Schneebeli-HermannPalaeoecology, Department of Physical Geography, Utrecht University, Budapestlaan 4, 3584 CD Utrecht, Netherlands

NITC

SITC

E>P

E>P

E<P

E<P

Land Mountain

Warm temperate humid Tropical humid

Subtropical arid

Tropical semi humid

Cold temperate Cool temperate

SouthChina

Tethys Ocean

Page 2: Extinguishing a Permian World - WordPress.com · are fusulinid foraminifera, rugose and tabulate corals, and the arthropod class Trilobita. The numerous hypotheses about the causes

288 GEOLOGY, March 2012

Permian macrofl ora (distribution of biomes) largely support the general climate zone distribution with strong seasonality for the Late Permian, as derived from climate models (Rees et al., 2002). The glob-ally less-differentiated Early Triassic fl ora were interpreted to indicate warm-temperate climate up to 70°N (Ziegler et al., 1993). Unfortunately, the time constraint of most macrofl oral records to infer detailed climate evolution is limited by the lack of calibration. Higher chronological reso-lution is achieved by implementing palynological data. Despite uncertain-ties regarding the botanical affi nities of 250 Ma spores and pollen, the approach of distinguishing sporomorphs according to the requirements of their parent plants with respect to water availability has been applied successfully to describe relative changes in humidity (e.g., Hochuli et al., 2010). In Norway, chemostratigraphically calibrated palynological records indicate a distinct shift from pollen-dominated (conifers and seed ferns) to spore-dominated (ferns and lycophytes) assemblages across the Perm-ian–Triassic transition. This change has been interpreted as a shift to more humid conditions in the earliest Triassic (Hochuli et al., 2010). Also, in the monsoon-dominated region of the southern subtropics, increasing spore abundances occur toward the Early Triassic (Hermann et al., 2012). Thus, the warming in the latest Permian is associated with changing global and regional precipitation and evaporation patterns resulting in more humid conditions in the northern mid-latitudes and southern subtropics.

The distribution and diversity of ammonoids in space and time, intro-duced as the latitudinal gradient of generic richness (LGGR), have also been interpreted to refl ect latitudinal SST gradients (Brayard et al., 2006). For the earliest Triassic (Griesbachian), the LGGR is very low, which has initially been interpreted as refl ecting a warm equable global climate cor-responding to fl at SST gradients. This supports the climate models pre-dicting a lower SST gradient with increased CO2 levels. However, in the earliest Triassic, this approach is limited because the diversity of ammo-noid communities was also affected by the extinction event.

Evidence for the modeled monsoonal circulation during Permian–Triassic times has been inferred from sedimentary records (e.g., Mutti and Weissert, 1995). Widespread redbed deposits are assumed to represent areas with seasonal rainfall (e.g., Parrish, 1993). For the Late Permian, a zone of high evaporation in the tropics can be delineated by mapping the global distribution of evaporites, reef carbonate, and coal deposits (Ziegler et al. 2003) (Fig. 1). Correlation of terrestrial sequences with marine suc-cessions is hampered by the different biostratigraphic frameworks, but nevertheless the change from paleosols with coals to paleosols contain-ing green–red-mottled claystones in Antarctica has been interpreted as a change from a dryer, cooler climate in the Late Permian to more humid, warmer conditions in the Early Triassic (Retallack and Krull, 1999).

In summary, the data of Joachimski et al. show that in the tropics, SST started to increase in the latest Permian prior to the main extinction event, continued increasing during the extinction into the earliest Trias-sic, and may have reached values of >32 °C. Additional paleotemperature records of comparable quality are now needed from less-condensed suc-cessions (i.e., with higher stratigraphic resolution) to enable us to further disentangle biotic and abiotic events around the Permian–Triassic bound-ary. Furthermore, paleotemperature data from other proxies may in turn strengthen the use of conodont paleothermometry in Permian to Triassic climate reconstructions.

REFERENCES CITEDBenton, M.J., Tverdokhlebov, V.P., and Surkov, M.V., 2004, Ecosystem remodel-

ling among vertebrates at the Permian–Triassic boundary in Russia: Nature, v. 432, p. 97–100, doi:10.1038/nature02950.

Barham, M., Joachimski, M.M., Murray, J., and Williams, D.M., 2012, Diagenetic alteration of the structure and δ18O signature of Palaeozoic fi sh and conodont apatite: Potential use for corrected isotope signatures in palaeoenvironmental interpretation: Chemical Geology, doi:10.1016/j.chemgeo.2011.12.026.

Brayard, A., Bucher, H., Escarguel, G., Fluteau, F., Bourquin, S., and Galfetti, T., 2006, The Early Triassic ammonoid recovery: Paleoclimatic signifi cance

of diversity gradients: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 239, p. 374–395, doi:10.1016/j.palaeo.2006.02.003.

Epstein, S., Buchsbaum, R., Lowenstam, H., and Urey, H.C., 1951, Carbonate-water isotopic temperature scale: Geological Society of America Bulletin, v. 62, no. 4, p. 417–426, doi:10.1130/0016-7606(1951)62[417:CITS]2.0.CO;2.

Hermann, E., Hochuli, P.A., Bucher, H., Brühwiler, T., Hautmann, M., Ware, D., Weissert, H., Roohi, G., and Yasseen, A., ur-Rehman, K., 2012, Climatic os-cillations at the onset of the Mesozoic inferred from palynological records from the North Indian Margin, Journal of the Geological Society, London, doi:10.1144/0016-76492010-130.

Hochuli, P.A., Vigran, J.O., Hermann, E., and Bucher, H., 2010, Multiple cli-matic changes around the Permian-Triassic boundary event revealed by an expanded palynological record from mid-Norway: Geological Society of America Bulletin, v. 122, p. 884–896, doi:10.1130/B26551.1.

Huang, C., Tong, J., Hinnov, L., and Chen, Z.Q., 2011, Did the great dying of life take 700 k.y.? Evidence from global astronomical correlation of the Permian-Triassic boundary interval: Geology, v. 39, no. 8, p. 779–782, doi:10.1130/G32126.1.

Joachimski, M.M., and Buggisch, W., 2002, Conodont apatite δ18O signatures in-dicate climatic cooling as a trigger of the Late Devonian mass extinction: Geology, v. 30, p. 711–714, doi:10.1130/0091-7613(2002)030<0711:CAOSIC>2.0.CO;2.

Joachimski, M.M., Lai, X., Shen, S., Jiang, H., Luo, G., Chen, B., Chen, J., and Sun, Y., 2012, Climate warming in the latest Permian and the Permian–Triassic mass extinction: Geology, v. 40, p. 195–198, doi:10.1130/G32707.1.

Kearsey, T., Twitchett, R.J., Price, G.D., and Grimes, S.T., 2009, Isotope excur-sions and palaeotemperature estimates from the Permian/Triassic boundary in the Southern Alps (Italy): Palaeogeography, Palaeoclimatology, Palaeo-ecology, v. 279, p. 29–40, doi:10.1016/j.palaeo.2009.04.015.

Kiehl, J.T., and Shields, C.A., 2005, Climate simulation of the latest Permian: Implications for mass extinction: Geology, v. 33, p. 757–760, doi:10.1130/G21654.1.

Kutzbach, J.E., and Ziegler, A.M., 1993, Simulation of Late Permian climate and biomes with an atmosphere–ocean model: comparisons with observa-tions: Philosophical Transactions of the Royal Society of London, Series B, v. 341, p. 327–340, doi:10.1098/rstb.1993.0118.

Mutti, M., and Weissert, H., 1995, Triassic monsoonal climate and its signature in Ladinian–Carnian carbonate platforms (Southern Alps, Italy): Journal of Sedimentary Research, v. B65, p. 357–367.

Parrish, J.T., 1993, Climate of the Supercontinent Pangea: The Journal of Geol-ogy, v. 101, p. 215–233, doi:10.1086/648217.

Raup, D.M., and Sepkoski, J.J., 1982, Mass extinctions in the marine fossil re-cord: Science, v. 215, p. 1501–1503, doi:10.1126/science.215.4539.1501.

Rees, P.M., Ziegler, A.M., Gibbs, M.T., Kutzbach, J.E., Behling, P.J., and Row-ley, D.B., 2002, Permian phytogeographic patterns and climate data/model comparisons: The Journal of Geology, v. 110, p. 1–31, doi:10.1086/324203.

Retallack, G.J., and Krull, E.S., 1999, Landscape ecological shift at the Permian–Triassic boundary in Antarctica: Australian Journal of Earth Sciences, v. 46, p. 785–812, doi:10.1046/j.1440-0952.1999.00745.x.

Roscher, M., Stordal, F., and Svensen, H., 2011, The effect of global warming and global cooling on the distribution of the latest Permian climate zones: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 309, p. 186–200, doi:10.1016/j.palaeo.2011.05.042.

Shen, S.Z., and 21 others, 2011, Calibrating the end-Permian mass extinction: Science, v. 334, p. 1367–1372, doi:10.1126/science.1213454.

Smith, A.G., Smith, D.G., and Funnell, B.M., 1994, Atlas of Mesozoic and Ceno-zoic Coastlines: Cambridge, UK, Cambridge University Press, 109 p.

Svensen, H., Planke, S., Polozov, A.G., Schmidbauer, N., Corfu, F., Podladchikov, Y.Y., and Jamtveit, B., 2009, Siberian gas venting and the end-Permian en-vironmental crisis: Earth and Planetary Science Letters, v. 277, p. 490–500, doi:10.1016/j.epsl.2008.11.015.

Urey, H.C., 1947, The thermodynamic properties of isotopic substances: Journal of the Chemical Society, v. 1947, p. 562–581.

Vennemann, T.W., Fricke, H.C., Blake, R.E., O’Neil, J.R., and Colman, A., 2002, Oxygen isotope analysis of phosphates: a comparison of techniques for analysis of Ag3PO4: Chemical Geology, v. 185, p. 321–336, doi:10.1016/S0009-2541(01)00413-2.

Ziegler, A.M., Eshel, G., Rees, P.M., Rothfus, T.A., Rowley, D.B., and Sunderlin, D., 2003, Tracing the tropics across land and sea: Permian to present: Le-thaia, v. 36, p. 227–254, doi:10.1080/00241160310004657.

Ziegler, A.M., Parrish, M., Yao, J., Gyllenhaal, E.D., Rowley, D.B., Parrish, J.T., Nie, S., Bekker, A., and Hulver, M.L., 1993, Early Mesozoic phytogeogra-phy and climate: Philosophical Transactions of the Royal Society of Lon-don, Series B, v. 341, p. 297–305, doi:10.1098/rstb.1993.0115.

Printed in USA