12.11.2013 Views

Deccan Volcanism Linked to the Cretaceous-Tertiary Boundary ...

Deccan Volcanism Linked to the Cretaceous-Tertiary Boundary ...

Deccan Volcanism Linked to the Cretaceous-Tertiary Boundary ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

JOURNAL GEOLOGICAL SOCIETY OF INDIA<br />

Vol.78, November 2011, pp.399-428<br />

<strong>Deccan</strong> <strong>Volcanism</strong> <strong>Linked</strong> <strong>to</strong> <strong>the</strong> <strong>Cretaceous</strong>-<strong>Tertiary</strong> <strong>Boundary</strong><br />

Mass Extinction: New Evidence from ONGC Wells in<br />

<strong>the</strong> Krishna-Godavari Basin<br />

G. KELLER 1 , P.K. BHOWMICK 2 , H. UPADHYAY 2 , A. DAVE 2 , A.N. REDDY 3 ,<br />

B.C. JAIPRAKASH 3 and T. ADATTE 4<br />

1 Geosciences Department, Prince<strong>to</strong>n University, Prince<strong>to</strong>n, NJ 08544, USA,<br />

2 KDMIPE, ONGC, Dehradun, India<br />

3 ONGC, Regional Geoscience Labora<strong>to</strong>ry, Chennai, India<br />

4 Geological and Paleon<strong>to</strong>logical Institute, Anthropole, CH-1015 Lausanne, Switzerland<br />

Email: gkeller@prince<strong>to</strong>n.edu<br />

Abstract: A scientific challenge is <strong>to</strong> assess <strong>the</strong> role of <strong>Deccan</strong> volcanism in <strong>the</strong> <strong>Cretaceous</strong>-<strong>Tertiary</strong> boundary (KTB)<br />

mass extinction. Here we report on <strong>the</strong> stratigraphy and biologic effects of <strong>Deccan</strong> volcanism in eleven deep wells from<br />

<strong>the</strong> Krishna-Godavari (K-G) Basin, Andhra Pradesh, India. In <strong>the</strong>se wells, two phases of <strong>Deccan</strong> volcanism record <strong>the</strong><br />

world’s largest and longest lava mega-flows interbedded in marine sediments in <strong>the</strong> K-G Basin about 1500 km from <strong>the</strong><br />

main <strong>Deccan</strong> volcanic province. The main phase-2 eruptions (~80% of <strong>to</strong>tal <strong>Deccan</strong> Traps) began in C29r and ended at<br />

or near <strong>the</strong> KTB, an interval that spans planktic foraminiferal zones CF1-CF2 and most of <strong>the</strong> nannofossil Micula prinsii<br />

zone, and is correlative with <strong>the</strong> rapid global warming and subsequent cooling near <strong>the</strong> end of <strong>the</strong> Maastrichtian. The<br />

mass extinction began in phase-2 preceding <strong>the</strong> first of four mega-flows. Planktic foraminifera suffered a 50% drop in<br />

species richness. Survivors suffered ano<strong>the</strong>r 50% drop after <strong>the</strong> first mega-flow, leaving just 7 <strong>to</strong> 8 survivor species. No<br />

recovery occurred between <strong>the</strong> next three mega-flows and <strong>the</strong> mass extinction was complete with <strong>the</strong> last phase-2 megaflow<br />

at <strong>the</strong> KTB. The mass extinction was likely <strong>the</strong> consequence of rapid and massive volcanic CO 2<br />

and SO 2<br />

gas<br />

emissions, leading <strong>to</strong> high continental wea<strong>the</strong>ring rates, global warming, cooling, acid rains, ocean acidification and a<br />

carbon crisis in <strong>the</strong> marine environment.<br />

<strong>Deccan</strong> volcanism phase-3 began in <strong>the</strong> early Danian near <strong>the</strong> C29R/C29n boundary correlative with <strong>the</strong> planktic<br />

foraminiferal zone P1a/P1b boundary and accounts for ~14% of <strong>the</strong> <strong>to</strong>tal volume of <strong>Deccan</strong> eruptions, including four of<br />

Earth’s longest and largest mega-flows. No major faunal changes are observed in <strong>the</strong> intertrappeans of zone P1b, which<br />

suggests that environmental conditions remained <strong>to</strong>lerable, volcanic eruptions were less intense and/or separated by<br />

longer time intervals thus preventing runaway effects. Alternatively, early Danian assemblages evolved in adaptation <strong>to</strong><br />

high-stress conditions in <strong>the</strong> aftermath of <strong>the</strong> mass extinction and <strong>the</strong>refore survived phase-3 volcanism. Full marine<br />

biotic recovery did not occur until after <strong>Deccan</strong> phase-3. These data suggest that <strong>the</strong> catastrophic effects of phase-2<br />

<strong>Deccan</strong> volcanism upon <strong>the</strong> <strong>Cretaceous</strong> planktic foraminifera were a function of both <strong>the</strong> rapid and massive volcanic<br />

eruptions and <strong>the</strong> highly specialized faunal assemblages prone <strong>to</strong> extinction in a changing environment. Data from <strong>the</strong><br />

K-G Basin indicates that <strong>Deccan</strong> phase-2 alone could have caused <strong>the</strong> KTB mass extinction and that impacts may have<br />

had secondary effects.<br />

Keywords: <strong>Cretaceous</strong>-<strong>Tertiary</strong>, Mass extinction, <strong>Deccan</strong> volcanism, Longest lava flows, Krishna-Godavari Basin.<br />

INTRODUCTION<br />

The biologic and environmental effects of <strong>Deccan</strong><br />

volcanism and its potential cause-and-effect relationship with<br />

<strong>the</strong> demise of <strong>the</strong> dinosaurs and <strong>the</strong> <strong>Cretaceous</strong>-<strong>Tertiary</strong><br />

boundary (KTB) mass extinction are <strong>the</strong> major unsolved<br />

problems in KTB studies <strong>to</strong>day. The <strong>Deccan</strong> volcanic<br />

province is one of <strong>the</strong> largest volcanic eruptions in Earth’s<br />

his<strong>to</strong>ry and <strong>to</strong>day covers an area of 512,000 km 2 (Fig. 1A),<br />

or about <strong>the</strong> size of France or Texas. The original size prior<br />

<strong>to</strong> erosion is estimated <strong>to</strong> have been 1.5 million km 2 and <strong>the</strong><br />

volume of lava extruded about 1.2 million km 3 , which <strong>to</strong>day<br />

can be seen as layers of lava flows with a <strong>to</strong>tal thickness of<br />

3500 m (Fig. 1B; Chenet et al. 2007).<br />

<strong>Deccan</strong> volcanism has been advocated as <strong>the</strong> potential<br />

0016-7622/2011-78-5-399/$ 1.00 © GEOL. SOC. INDIA


400 G. KELLER AND OTHERS<br />

Fig.1. (A) Map of India with current distribution of <strong>Deccan</strong> traps, including <strong>the</strong> Earth’s longest lava flows <strong>to</strong> <strong>the</strong> Krishna-Godavari<br />

Basin and out in<strong>to</strong> <strong>the</strong> Bay of Bengal; black dots mark KTB locations studied. (B) Layered <strong>Deccan</strong> lava flows (traps) form<br />

Mahalabeshwar. (C) Map of <strong>the</strong> K-G Basin with locations of Rajahmundry quarries and ONGC wells studied for this report.<br />

(D) Lower and upper basalt flows exposed in <strong>the</strong> Gauripatnam quarry of Rajahmundry separated by intertrappean sediments of<br />

earliest Danian (zone P1a) age.<br />

cause for <strong>the</strong> KTB catastrophe for over thirty years (e.g.,<br />

McLean, 1978, 1985; Courtillot et al. 1986, 1988; Duncan<br />

and Pyle, 1988; Venkatesan et al. 1993; Raju et al. 1995;<br />

Sheth et al. 2001; Pande et al. 2004). But this hypo<strong>the</strong>sis<br />

was considered unlikely because a direct link <strong>to</strong> <strong>the</strong> mass<br />

extinction remained elusive in <strong>the</strong> absence of <strong>Deccan</strong> lava<br />

flows interbedded with marine sediments rich in microfossils<br />

<strong>to</strong> assess <strong>the</strong> nature of <strong>the</strong> mass extinction, and also because<br />

volcanism was generally believed <strong>to</strong> have occurred over at<br />

least one million years, leaving sufficient time for recovery<br />

between eruptions.<br />

Over <strong>the</strong> past several years a number of multidisciplinary<br />

studies have changed this perception and<br />

directly linked <strong>Deccan</strong> volcanism <strong>to</strong> <strong>the</strong> KTB mass<br />

extinction: (1) Improved dating of <strong>the</strong> 3500 m thick <strong>Deccan</strong><br />

lava pile revealed that <strong>the</strong> major eruptions occurred in three<br />

phases with <strong>the</strong> initial eruption, phase-1, in <strong>the</strong> late<br />

Maastrichtian base of C30n (67.4 Ma), <strong>the</strong> main phase-2 in<br />

C29r below <strong>the</strong> KTB, and <strong>the</strong> last phase-3 in <strong>the</strong> early Danian<br />

base C29n (Fig. 2; Chenet et al. 2007, 2008, 2009; Jay and<br />

Widdowson, 2008). (2) Massive eruptions created <strong>the</strong><br />

longest and largest lava flows on Earth (Self et al. 2008a,<br />

b). And (3) a direct link between <strong>the</strong> main phase of <strong>Deccan</strong><br />

eruptions and <strong>the</strong> KTB mass extinction was documented in<br />

Rajahmundry quarries (Andhra Pradesh) and Jhilmili<br />

(Madhya Pradesh) based on planktic foraminifera, which<br />

suffered <strong>the</strong> most devastating mass extinction globally<br />

(Keller et al. 2008, 2009a, b, c). At <strong>the</strong>se two localities, no<br />

Maastrichtian marine sediments are exposed but <strong>the</strong> earliest<br />

Danian species are present in <strong>the</strong> intertrappean sediments<br />

between phase-2 and phase-3 mega-flows.<br />

Still missing from <strong>the</strong>se early results is critical<br />

JOUR.GEOL.SOC.INDIA, VOL.78, NOV. 2011


DECCAN VOLCANISM LINKED TO THE CRETACEOUS-TERTIARY BOUNDARY MASS EXTINCTION 401<br />

Fig.2. Relative thickness of <strong>Deccan</strong> lava flows in each of <strong>the</strong> three<br />

phases of volcanic eruptions calculated as percent of <strong>to</strong>tal<br />

<strong>Deccan</strong> trap thickness. Ages based on paleomagnetic time<br />

scale (modified from Chenet et al. 2007, 2008).<br />

information concerning <strong>the</strong> onset and age of <strong>the</strong> main <strong>Deccan</strong><br />

phase, <strong>the</strong> nature and tempo of <strong>the</strong> mass extinction relative<br />

<strong>to</strong> eruption pulses, and <strong>the</strong> number of <strong>the</strong> longest lava flows,<br />

here termed ‘mega-flows’. No outcrops exist that can yield<br />

this information because deposition in India occurred mainly<br />

in terrestrial and some shallow inner neritic environments<br />

(e.g., Rajahmundry, Jhilmili), which lack diverse fossil<br />

assemblages, and <strong>the</strong> lava mega-flows are fused, preventing<br />

environmental studies (Fig. 1D).<br />

The only area that can yield <strong>the</strong> missing information is<br />

in <strong>the</strong> Krishna-Godavari Basin, which spans about 75 km<br />

from Rajahmundry <strong>to</strong>wards <strong>the</strong> Bay of Bengal (Fig. 1C).<br />

During KTB time, <strong>the</strong>re was progressive deepening seaward<br />

from <strong>the</strong> inner neritic environment of Rajahmundry <strong>to</strong> outer<br />

neritic depths (100-150 m) accompanied by increasingly<br />

diverse planktic foraminiferal assemblages (Jaiprakash et<br />

al. 1993; Raju et al. 1994). During <strong>the</strong> Cenozoic, <strong>the</strong> high<br />

sediment input by <strong>the</strong> Krishna and Godavari drainage<br />

systems and a complex system of horst and graben structures<br />

resulted in rapid subsidence and <strong>to</strong>day <strong>the</strong> KTB sequences<br />

dip between 2500 m <strong>to</strong> 3500 m below <strong>the</strong> surface <strong>to</strong>wards<br />

<strong>the</strong> Bay of Bengal. The only access <strong>to</strong> <strong>the</strong>se critical KTB<br />

records are deep wells drilled by India’s Oil and Natural<br />

Gas Corporation Ltd. (ONGC) in <strong>the</strong> K-G Basin. The nature<br />

of information that can be gained from <strong>the</strong>se ONGC wells<br />

is evident from various publications by ONGC scientists<br />

that show a series of lava flows and intertrappean sediments<br />

spanning from <strong>the</strong> upper Maastrichtian <strong>to</strong> <strong>the</strong> lower<br />

Paleocene (e.g., Govindan, 1981; Mehrotra and Sargeant,<br />

1987; Raju et al. 1994, 1995, 1996; Jaiprakash et al. 1993;<br />

Prasad and Pundeer, 2002; Misra, 2005; Raju, 2008).<br />

Despite this effort, locating <strong>the</strong> precise position of <strong>the</strong> KTB<br />

in <strong>the</strong>se wells relative <strong>to</strong> <strong>the</strong> <strong>Deccan</strong> mega-flows remained<br />

elusive.<br />

With new insights gained from recent studies, reinvestigation<br />

of ONGC deep wells yielded critical<br />

information on <strong>the</strong> relative timing of <strong>the</strong> largest and longest<br />

lava mega-flows and <strong>the</strong> environmental conditions as<br />

inferred from planktic foraminiferal assemblages. Here we<br />

report <strong>the</strong> results from ten wells from <strong>the</strong> onshore K-G Basin<br />

and one offshore well (Fig. 1C). The results demonstrate:<br />

(a) up <strong>to</strong> four of Earth’s longest and largest lava mega-flows<br />

(traps) separated by intertrappean sediments in each phase-<br />

2 and phase-3; (b) <strong>the</strong> nature and tempo of <strong>the</strong> mass<br />

extinction, as indicated by sediments below <strong>the</strong> mega-flows<br />

(infratrappean), and between mega-flows (intertrappean);<br />

and c) <strong>the</strong> close link between <strong>Deccan</strong> volcanism and <strong>the</strong><br />

KTB mass extinction.<br />

BACKGROUND<br />

<strong>Deccan</strong> Eruption Phases<br />

Detailed paleomagnetic and radiometric dating ( 40 K/ 40 Ar<br />

and 40 Ar/ 39 Ar) of <strong>the</strong> 3500 m thick <strong>Deccan</strong> lava pile in <strong>the</strong><br />

western Ghats revealed that volcanic eruptions occurred in<br />

a series of rapid pulses grouped in three main phases (Chenet<br />

et al. 2007, 2008, 2009). A rough estimate of <strong>the</strong> relative<br />

volume can be estimated by <strong>the</strong> thickness of <strong>the</strong> major lava<br />

flows in each of <strong>the</strong> three eruption phases across <strong>the</strong> <strong>Deccan</strong><br />

volcanic province (Fig. 2). Syed F.R. Khadri estimated <strong>the</strong><br />

cumulative maximum <strong>to</strong>tal thickness as possibly reaching<br />

5000 m (personal communication, 2011). Based on <strong>the</strong> more<br />

conservative estimate of 3500 m, about 6% of this lava pile<br />

is attributed <strong>to</strong> <strong>the</strong> initial and smallest phase-1 (Latifwadi<br />

Formation), which is dated at ~67.4 Ma near <strong>the</strong> base of<br />

C30n. Less intense volcanic eruptions likely continued<br />

during <strong>the</strong> late Maastrichtian.<br />

The main phase-2 (Jawhar and Ambenali Formations)<br />

occurred in C29r below <strong>the</strong> KTB and accounts for about<br />

~80% of <strong>the</strong> <strong>to</strong>tal <strong>Deccan</strong> thickness. The shear volume of<br />

phase-2 suggests that this eruption phase could have been<br />

detrimental <strong>to</strong> life. Chenet et al. (2007, 2008) estimated that<br />

each of 30 <strong>to</strong> 100 major eruptive pulses in phase-2 emitted<br />

volumes ranging from 20,000 km 3 <strong>to</strong> 120,000 km 3 , attained<br />

a thicknesses up <strong>to</strong> 200 m, and was emplaced over hundreds<br />

of kilometers in a relatively short time interval in C29r below<br />

JOUR.GEOL.SOC.INDIA, VOL.78, NOV. 2011


402 G. KELLER AND OTHERS<br />

<strong>the</strong> KTB. The last phase-3 (Mahalabeshwar Formation) is<br />

estimated at about 14% of <strong>the</strong> <strong>to</strong>tal lava pile and began in<br />

<strong>the</strong> early Danian at or near <strong>the</strong> base of C29n about 270 ky<br />

after <strong>the</strong> KTB mass extinction (Fig. 2; time scale of Cande<br />

and Kent, 1995).<br />

The Longest Lava Flows<br />

Paleomagnetic and geochemical studies have correlated<br />

<strong>the</strong> lower and upper basalt flows of <strong>the</strong> Rajahmundry quarries<br />

(Andhra Pradesh, Figs. 1D) <strong>to</strong> <strong>the</strong> Ambenali and<br />

Mahalabeshwar Formations, or phase-2 and phase-3 of <strong>the</strong><br />

<strong>Deccan</strong> lava pile, respectively (e.g., Knight et al. 2003, 2005;<br />

Baksi et al. 1994; Baksi, 2005; Jay and Widdowson, 2008;<br />

Jay et al. 2009). Age determinations based on 40 K/ 40 Ar and<br />

40 Ar/ 39 Ar place <strong>the</strong>se lava mega-flows in C29r and near <strong>the</strong><br />

base of C29n, respectively, although error margins are 1%<br />

or 0.6 Ma (e.g., Knight et al. 2003, 2005; Baksi, 2005).<br />

This marks <strong>the</strong> phase-2 and phase-3 eruptions as <strong>the</strong> largest<br />

and longest lava flows on Earth, spanning over 1500 km<br />

from <strong>the</strong> main <strong>Deccan</strong> province across India <strong>to</strong> <strong>the</strong> Krishna-<br />

Godavari Basin and out in<strong>to</strong> <strong>the</strong> Bay of Bengal (Figs. 1A,<br />

C). Self et al. (2008a) interpreted <strong>the</strong>se mega-flows as very<br />

large volume pahoehoe flow fields that for <strong>the</strong> last 400 km<br />

of <strong>the</strong>ir journey were confined <strong>to</strong> <strong>the</strong> pre-existing Godavari<br />

valley drainage system that channeled <strong>the</strong>ir flow <strong>to</strong> <strong>the</strong><br />

eastern estuaries near Rajahmundry and in<strong>to</strong> <strong>the</strong> Bay of<br />

Bengal. They estimated <strong>the</strong>se mega-flows as <strong>the</strong> world’s<br />

largest at 5000 km 3 of eruptive volume.<br />

<strong>Deccan</strong> and <strong>the</strong> KTB Mass Extinction<br />

The above studies demonstrate that <strong>Deccan</strong> volcanic<br />

phase-2 and phase-3 bracket <strong>the</strong> KTB mass extinction, but<br />

cannot establish a direct link. Paleon<strong>to</strong>logical investigations<br />

first discovered this link in 4 <strong>to</strong> 9 m thick intertrappeans<br />

between phase-2 and phase-3 mega-flows in six<br />

Rajahmundry basalt quarries (Figs.1D; Keller et al. 2008;<br />

Malarkodi et al. 2010). In this estuarine environment early<br />

Danian (zone P1a) planktic foraminiferal assemblages<br />

directly overlie <strong>the</strong> <strong>to</strong>p of phase-2 eruptions and indicate<br />

that <strong>the</strong> mass extinction occurred at or near <strong>the</strong> end of this<br />

volcanic phase (Keller et al. 2008). These results were<br />

confirmed in intertrappean sediments in central India<br />

(Jhilmili, Madhya Pradesh, Keller et al. 2009a, b, c).<br />

MATERIAL AND METHODS<br />

ONGC wells from <strong>the</strong> Krishna-Godavari Basin were<br />

chosen for <strong>the</strong>ir apparent continuity across <strong>the</strong> KTB<br />

transition, <strong>the</strong> number of basalt flows and intertrappean<br />

sediments, and <strong>the</strong> availability of recovered cores in addition<br />

<strong>to</strong> well cuttings. A <strong>to</strong>tal of eleven wells have been analyzed<br />

for this study, including one offshore well (G-4-6; Fig. 1C).<br />

Sample material from each well was collected in <strong>the</strong> ONGC<br />

core library in Rajahmundry and supplemented by samples<br />

from <strong>the</strong> core libraries at Dehradun and Chennai. Well<br />

samples were taken at 5 m or 1 m intervals from cuttings,<br />

and at 20 cm intervals in recovered core sections. About<br />

200 gr sediments were collected per sample for analysis<br />

and a <strong>to</strong>tal of 665 samples were analyzed. Samples were<br />

processed by standard micropaleon<strong>to</strong>logical techniques<br />

(Keller et al. 1995).<br />

For biostratigraphic analysis <strong>the</strong> washed residues of each<br />

sample was searched for foraminifera in several size<br />

fractions, including 38-63 µm, 63-105 µm, 105-150 µm,<br />

150-250 µm and >250 µm in order <strong>to</strong> analyze all species<br />

from <strong>the</strong> very small <strong>to</strong> <strong>the</strong> very large. Foraminifera were<br />

picked from each size fraction, identified at <strong>the</strong> species level<br />

and recorded. The species census data was filtered <strong>to</strong> exclude<br />

down-core contamination that is common in well cuttings.<br />

Theoretically, this means that any species first appearances<br />

may be <strong>the</strong> result of down-core contamination. To avoid<br />

such artificial range extension, we attributed any isolated<br />

species occurrences at <strong>the</strong> base of <strong>the</strong> range <strong>to</strong> down-core<br />

contamination. Last appearances of species were relied<br />

upon for stratigraphic age control, except for isolated<br />

species that could be <strong>the</strong> result of reworking. Based on this<br />

filtered dataset, combined with <strong>the</strong> stratigraphic control<br />

from core intervals, and <strong>the</strong> resistivity and gamma ray well<br />

log data, good age control was achieved for all wells<br />

analyzed.<br />

Preservation of foraminiferal tests in upper Maastrichtian<br />

sediments ranges from good <strong>to</strong> poor, with predominantly<br />

poor preservation in very deep wells (> 3000 m) and poor<br />

preservation due <strong>to</strong> dissolution effects in intertrappean<br />

sediments. In contrast, Danian assemblages are relatively<br />

well preserved. SEM’s of late Maastrichtian and early<br />

Danian planktic foraminifera illustrate faunal assemblages<br />

of <strong>the</strong> KTB transition in <strong>the</strong> Krishna-Godavari Basin.<br />

STRATIGRAPHY OF DECCAN BASALT<br />

MEGA-FLOWS<br />

The lithostratigraphy of <strong>the</strong> Krishna-Godavari Basin as<br />

determined from outcrops (Rajahmundry area) and<br />

subsurface wells is shown in Fig.3. In surface outcrops, <strong>the</strong><br />

Rajahmundry sands<strong>to</strong>ne overlies <strong>the</strong> upper <strong>Deccan</strong> megaflows,<br />

or phase-3, and <strong>the</strong> Tirupati sands<strong>to</strong>ne underlies <strong>the</strong><br />

lower <strong>Deccan</strong> mega-flows, or phase-2. We observed in both<br />

outcrops and recent drilling that <strong>the</strong> Tirupati sands<strong>to</strong>ne ends<br />

with a marine bed consisting of abundant Turritella, few<br />

JOUR.GEOL.SOC.INDIA, VOL.78, NOV. 2011


DECCAN VOLCANISM LINKED TO THE CRETACEOUS-TERTIARY BOUNDARY MASS EXTINCTION 403<br />

Fig.3. Generalized lithostratigraphy of <strong>the</strong> Krishna-Godavari Basin (modified after<br />

Venkatarengan et al. 1993). Note <strong>the</strong> outcrop lithostratigraphy is representative of<br />

Rajahmundry quarries. Well locations studied for this report are located in both sand<br />

(landward) and shale lithologies (basinward).<br />

o<strong>the</strong>r macrofossils and rare benthic foraminifera. Below <strong>the</strong><br />

Turritella bed sediments consist of clay and silt, representing<br />

flood plain and paleosoils, which alternate with sands<strong>to</strong>nes<br />

indicative of fluvial and channel environments. At <strong>the</strong> <strong>to</strong>p<br />

of phase-3 mega-flows, an erosion surface marks a major<br />

hiatus, which is overlain by sands<strong>to</strong>ne of middle Oligocene<br />

age and corresponding <strong>to</strong> <strong>the</strong> Nimmakuru sands<strong>to</strong>ne (Fig.3).<br />

Phase-2 and phase-3 mega-flows are separated by<br />

intertrappean sediments consisting of dolomitic muds<strong>to</strong>ne<br />

at <strong>the</strong> base followed by estuarine <strong>to</strong> shallow marine limes<strong>to</strong>ne<br />

and paleosoil in <strong>the</strong> upper part (Keller et al. 2008).<br />

Basinwards from Rajahmundry, <strong>the</strong> sands<strong>to</strong>nes, shales<br />

and limes<strong>to</strong>nes thin out and are replaced by marly and clayey<br />

shales (Fig. 3). Wells for this study were taken from<br />

middle/outer neritic environments that cut through <strong>the</strong> distal<br />

end-members of <strong>the</strong> Tirupati sands<strong>to</strong>ne, Razole Formation<br />

and Palakollu shale, as well as from deeper neritic<br />

environments dominated by shales (Fig. 3). In most of <strong>the</strong><br />

wells, <strong>the</strong> Razole Formation consists of up <strong>to</strong> eight and<br />

occasionally nine <strong>Deccan</strong> mega-flows separated by<br />

intertrappeans. Three <strong>to</strong> four and occasionally five megaflows<br />

are located below <strong>the</strong> KTB and mark <strong>Deccan</strong> phase-<br />

2 and three <strong>to</strong> four mega-flows mark <strong>Deccan</strong> phase-3 in <strong>the</strong><br />

lower Danian (Jaiprakash et al. 1993; Raju et al. 1994, 1995,<br />

1996; Misra, 2005; Raju, 2008; Jay and Widdowson, 2008;<br />

this study).<br />

In <strong>the</strong> Rajahmundry quarries, <strong>the</strong> pulsed volcanic<br />

eruptions that created <strong>the</strong> mega-flows of phase-2 in <strong>the</strong><br />

Krishna-Godavari Basin can be differentiated (Keller et al.<br />

2008), but are not separated by intertrappeans as <strong>the</strong>y are in<br />

<strong>the</strong> Krishna-Godavari Basin wells and logs (Fig. 4).<br />

JOUR.GEOL.SOC.INDIA, VOL.78, NOV. 2011<br />

Resistivity values for normal<br />

sands<strong>to</strong>nes, silts<strong>to</strong>nes and shales vary<br />

between 1-6 qm, unless <strong>the</strong>y are<br />

hydrocarbon bearing. In contrast,<br />

resistivity values for <strong>the</strong> basalts vary<br />

anywhere from 50 qm <strong>to</strong> > 200 qm<br />

depending on <strong>the</strong> degree of sediments<br />

incorporated as a result of erosion at<br />

<strong>the</strong> base and <strong>to</strong>p. In all wells, e – logs<br />

show high <strong>to</strong> very high resistivity<br />

peaks against <strong>the</strong> basalt flows. Most<br />

basalts are seen as a distinct resistivity<br />

peak on <strong>the</strong> logs, whereas some<br />

basalts are less distinct due <strong>to</strong> mixed<br />

basalt and sediments. Gamma logs<br />

show significantly lower values in<br />

basalt flows relative <strong>to</strong> <strong>the</strong> intertrappean<br />

clastic sediments (Fig. 4).<br />

The cores and drill cuttings from<br />

<strong>the</strong>se mega-flows are dark grey <strong>to</strong> greenish grey in colour,<br />

very hard and compact.<br />

<strong>Deccan</strong> mega-flows in <strong>the</strong> Krishna-Godavari Basin are<br />

generally 5 m <strong>to</strong> 15 m thick, except for two wells where <strong>the</strong><br />

lava flows in phase-3 are 60 m thick (Fig. 5, CTP-A and<br />

RZL-A). The variation in <strong>the</strong> number of basalt flows and<br />

variable thicknesses can be explained by <strong>to</strong>pography and<br />

erosion. The variable thickness of intertrappeans is a function<br />

of <strong>to</strong>pography, subsidence and influx of sediments eroded<br />

from shallower areas, as indicated by an abundance of sand<br />

in some wells (Figs. 3, 4, e.g., MTP-A, CTP-A, RZL-1,<br />

PNM-A). The overall pattern of <strong>the</strong> mega-flows in each<br />

volcanic phase is consistent with sheet flows of very large<br />

volume pahoehoe flow fields, as suggested by Self et al.<br />

(2008a).<br />

BIOSTRATIGRAPHY<br />

Identifying <strong>the</strong> KT <strong>Boundary</strong><br />

The high-resolution biostratigraphic zonal scheme<br />

developed by Keller et al. (1995, 2002) for <strong>the</strong> Danian and<br />

by Li and Keller (1998a, b) for <strong>the</strong> Maastrichtian (Fig. 5) is<br />

applied in <strong>the</strong> K-G Basin wells. Ages for biozones are<br />

calculated for <strong>the</strong> time scales of Cande and Kent (1995;<br />

KTB at 65 Ma) and Gradstein et al. (2004; KTB at 65.5<br />

Ma, Fig. 6). The KTB can be globally identified based on a<br />

number of criteria: <strong>the</strong> mass extinction of planktic<br />

foraminifera, <strong>the</strong> evolution of <strong>the</strong> first Danian species within<br />

a few centimeters above <strong>the</strong> extinction horizon, a clay layer<br />

with a millimeter thin oxidized red layer at <strong>the</strong> base, an<br />

iridium anomaly in <strong>the</strong> red layer, and a 2 <strong>to</strong> 3 permil δ 13 C


404 G. KELLER AND OTHERS<br />

Fig.4. Correlation of Krishna-Godavari Basin wells based on biostratigraphy, <strong>Deccan</strong> mega-flows and well log data (gamma and resistivity). Note that <strong>Deccan</strong> phase-2 and phase-3<br />

each contain three <strong>to</strong> four lava flows that represent Earth’s largest longest mega-flows separated by intertrappean sediments. These phase-2 and phase-3 mega-flows correlate<br />

with <strong>the</strong> “fused” lower and upper traps in Rajahmundry quarries.<br />

JOUR.GEOL.SOC.INDIA, VOL.78, NOV. 2011


DECCAN VOLCANISM LINKED TO THE CRETACEOUS-TERTIARY BOUNDARY MASS EXTINCTION 405<br />

Fig.5. Biostratigraphy and planktic foraminiferal zonal scheme by Keller et al. (1995, 2002) and Li and Keller (1998a,b) for <strong>the</strong> late<br />

Maastrichtian and early Danian in <strong>the</strong> El Kef stra<strong>to</strong>type section and point (GSSP) with illustrations of index species, <strong>the</strong> Ir<br />

anomaly (Rocchia et al. 1996) and δ 13 C shift (Keller and Lindinger, l989) are correlated with <strong>the</strong> paleomagnetic time scale and<br />

<strong>Deccan</strong> volcanic events. The nannofossil zonal scheme (*) by Tantawy (2003), and <strong>the</strong> planktic foraminiferal zonation (**) by<br />

Bergren et al. (1995) are shown for comparison.<br />

shift across <strong>the</strong> clay layer (Fig. 5; Keller et al. l995; Cowie<br />

et al. 1989; Remane et al. 1999). Among <strong>the</strong>se, only<br />

extinction and evolution events are unique KTB-defining<br />

criteria. All o<strong>the</strong>rs (Ir anomaly, clay and red layers, δ 13 C<br />

shift) are KTB-supporting criteria that cannot stand alone<br />

as KTB markers because <strong>the</strong>y are not unique events. In India,<br />

all of <strong>the</strong>se KTB markers are present in <strong>the</strong> Meghalaya<br />

section (Gertsch et al. 2011), but <strong>to</strong> date <strong>the</strong> Ir anomaly and<br />

δ 13 C shift have not been documented in <strong>the</strong> Krishna-<br />

Godavari Basin due <strong>to</strong> lack of suitable marine sequences<br />

with high-resolution sample spacing.<br />

Rajahmundry Quarries<br />

A close link between <strong>Deccan</strong> volcanism and <strong>the</strong> KTB<br />

mass extinction was first established in Rajahmundry<br />

quarries based on zone P1a planktic foraminiferal<br />

assemblages in intertrappean sediments between phase-2<br />

and phase-3 lava mega-flows (Fig. 1D; Keller et al. 2008).<br />

These intertrappeans contain <strong>the</strong> first Danian species, which<br />

evolved in <strong>the</strong> aftermath of <strong>the</strong> mass extinction (e.g.,<br />

Parvularugoglobigerina extensa, P. eugubina, Globoconusa<br />

daubjergensis, Woodringina horners<strong>to</strong>wnensis; Fig. 5). The<br />

index species P. eugubina marks this assemblage as zone<br />

P1a, which <strong>to</strong>ge<strong>the</strong>r with zone P0 (boundary clay) spans<br />

C29r above <strong>the</strong> KTB. The boundary clay was not observed.<br />

Based on this information, <strong>the</strong> KTB was placed at <strong>the</strong> base<br />

of <strong>the</strong> intertrappean above phase-2 (Figs. 1D). No sediments<br />

are present above <strong>the</strong> phase-3 mega-flows and sediments<br />

below phase-2 consist of beach sand with rare benthic<br />

foraminifera.<br />

Krishna-Godavari Basin<br />

Maastrichtian below phase-2 mega-flows: Zones CF2-CF3<br />

In <strong>the</strong> Krishna-Godavari Basin, about 75 km seaward<br />

from Rajahmundry, sediment deposition occurred in a<br />

shallow middle neritic environment (


406 G. KELLER AND OTHERS<br />

Fig.6. Late Maastrichtian and early Danian planktic foraminiferal zonal schemes and ages of biozones based on two time scales with <strong>the</strong><br />

KT boundary at 65.0 and 65.5 Ma. Also shown for comparison are <strong>the</strong> zonal schemes of Caron (1985), Berggren et al. (1995) and<br />

<strong>the</strong> nannofossil zonation of Tantawy (2003). CF=<strong>Cretaceous</strong> Foraminifera.<br />

environments (Keller and Abramovich, 2009). In general,<br />

<strong>the</strong> most diverse assemblages were observed in <strong>the</strong><br />

cored intervals, such as shown for PLK-A and NSP-A wells<br />

(Figs. 7 and 8). Since <strong>the</strong> cored intervals contain <strong>the</strong> most<br />

reliable data, <strong>the</strong> lower diversity in core cuttings is likely an<br />

artifact of preservation and culling of species suspected <strong>to</strong><br />

be down-core contaminants. In wells with predominantly<br />

sand and silt deposition, such as MTP-A and PNM-A,<br />

(Figs. 12-13), foraminifera are rare due <strong>to</strong> increased<br />

dissolution and dilution by high sediment influx.<br />

Biostratigraphic control in <strong>the</strong>se wells is more difficult <strong>to</strong><br />

establish.<br />

In ten out of eleven wells analyzed for this study, late<br />

Maastrichtian faunal assemblages below phase-2 megaflows<br />

are typical of zones CF2-CF3. In <strong>the</strong> absence of<br />

Gansserina gansseri <strong>the</strong>se two zones cannot be<br />

differentiated. Zone CF2 spans 120 ky (65.66-65.78 Ma)<br />

and zone CF3 spans 1.21 Ma (65.78 <strong>to</strong> 66.99 Ma, time scale<br />

of Gradstein et al. 2004; Fig. 6). Zone CF3 is nearly<br />

equivalent <strong>to</strong> <strong>the</strong> Micula murus nannofossil zone. Saxena<br />

and Misra (1994, 1995) and von Salis and Saxena (1998)<br />

identified <strong>the</strong> M. murus zone assemblage below phase-2<br />

mega-flows in <strong>the</strong> RZL-A, NSP-B and PLK-A wells (Figs.<br />

7, 9, 12). Only in <strong>the</strong> offshore well G-4-F <strong>to</strong> <strong>the</strong> nor<strong>the</strong>ast of<br />

<strong>the</strong> K-G Basin (Fig. 1B) is <strong>the</strong> upper Maastrichtian missing<br />

(zones CF1-CF5, Fig. 14). In this well, a diverse assemblage<br />

of 43 species identifies zone CF6 (69.08-69.61 Ma)<br />

including <strong>the</strong> index species (first appearance of<br />

Contusotruncna contusa and last appearance of<br />

Globotruncana linneiana) underlying one mega-flow and<br />

<strong>the</strong> KTB. Zone CF6 precedes <strong>Deccan</strong> phase-1 dated about<br />

67.4 Ma. Late Maastrichtian planktic foraminiferal<br />

assemblages from <strong>the</strong> K-G Basin are illustrated in Plates<br />

1-3.<br />

Volcanic phase-2 intertrappeans: Zone CF1<br />

In <strong>the</strong> K-G wells, phase-2 intertrappeans are generally<br />

between 5 m and 15 m thick. But in two wells <strong>the</strong>y reach a<br />

maximum of 60 m (CTP-A and RZL-A; Figs. 1C, 11),<br />

probably due <strong>to</strong> <strong>to</strong>pographic lows and high sedimentary<br />

influx. Alternatively, this area may be cut by faults and<br />

thrusts. Paleomagnetic and radiometric analyses placed <strong>the</strong><br />

Rajahmundry lower traps (equivalent <strong>to</strong> phase-2 megaflows)<br />

in C29r below <strong>the</strong> KTB (e.g., Knight et al. 2003,<br />

2005; Baksi, 2005; Chenet et al. 2007, 2008, 2009), which<br />

is equivalent <strong>to</strong> zones CF2 (65.66-65.78 Ma) and CF1<br />

JOUR.GEOL.SOC.INDIA, VOL.78, NOV. 2011


DECCAN VOLCANISM LINKED TO THE CRETACEOUS-TERTIARY BOUNDARY MASS EXTINCTION 407<br />

Fig.7. ONGC well PLK-A with biostratigraphy, species occurrences, phase-2 and phase-3 mega-flows plotted against lithostratigraphy<br />

and e-logs (gamma and resistivity). Core segment shown is from <strong>the</strong> base of <strong>the</strong> section, and ano<strong>the</strong>r core segment from <strong>the</strong> last<br />

mega-flow in phase-3.<br />

Fig.8. ONGC well NSP-A with biostratigraphy, species occurrences, and phase-2 mega-flows plotted against lithostratigraphy and e-<br />

logs (gamma and resistivity). A 4 m cored interval below <strong>the</strong> first mega-flow of phase-2 records a 50% diversity crash, which<br />

appears <strong>to</strong> be related <strong>to</strong> <strong>the</strong> onset of <strong>Deccan</strong> phase-2 volcanism. No samples are available from <strong>the</strong> intertrappean of this well, but<br />

in o<strong>the</strong>r wells <strong>the</strong>re is ano<strong>the</strong>r 50% crash after <strong>the</strong> first mega-flow (see Fig. 7). Phase-3 mega-flows are not present in NSP-A. A<br />

cosmic spherule was found at 3365 m.<br />

JOUR.GEOL.SOC.INDIA, VOL.78, NOV. 2011


408 G. KELLER AND OTHERS<br />

Fig.9. ONGC well NSP-B with biostratigraphy, species occurrences, phase-2 and phase-3 lava flows plotted against lithostratigraphy<br />

and e-logs (gamma and resistivity). Note <strong>the</strong> low species diversity in infra- and intertrappeans of phase-2 volcanism is partly due<br />

<strong>to</strong> poor preservation and dissolution effects.<br />

(65.50-65.66 Ma) and most of <strong>the</strong> nannofossil Micula prinsii<br />

zone (Figs. 5 and 6). Based on this correlation <strong>the</strong> phase-2<br />

mega-flows and intertrappeans of <strong>the</strong> Krishna-Godavari<br />

Basin were most likely deposited in zone CF1 with <strong>the</strong> onset<br />

of phase-2 volcanism in zone CF2 (Fig. 6). In <strong>the</strong> K-G Basin<br />

<strong>the</strong> fragile zone CF1 index species, Plummerita<br />

hantkeninoides, was not observed, ei<strong>the</strong>r because of poorly<br />

preserved assemblages, or because of exclusion due <strong>to</strong> stress<br />

conditions. Very few species are present in intertrappean<br />

sediments due <strong>to</strong> extinctions, dissolution and poor<br />

preservation due <strong>to</strong> acid rains associated with <strong>Deccan</strong><br />

volcanism. The stress conditions during deposition of <strong>the</strong>se<br />

intertrappeans are discussed below.<br />

KT boundary and early Danian: Zone P1a<br />

In Krishna-Godavari Basin wells, <strong>the</strong> KT boundary is<br />

identified in <strong>the</strong> intertrappean sediments between phase-2<br />

and phase-3 mega-flows by <strong>the</strong> lower Danian zone P1a<br />

planktic foraminiferal assemblages. Most wells contain<br />

species that evolved in <strong>the</strong> lower part of zone P1a (subzone<br />

P1a(1), including Parvularugogloigerina extensa,<br />

P. eugubina, Eoglobigerina edita, Globoconusa<br />

daubjergensis, G. taurica, and Woodringina horners<strong>to</strong>wnensis<br />

(Figs. 7, 11 and 12). However, most wells also<br />

contain species that evolved in <strong>the</strong> upper part of zone P1a<br />

(subzone P1a(2), such as Subbotina triloculinoides,<br />

Parasubbotina pseudobulloides, Globigerina pentagona,<br />

Chiloguembelina midwayensis, and Globanomalina<br />

archeocompressa (Plates 4, 5). Because of poor sample<br />

control, <strong>the</strong>se subzones cannot be identified at this time and<br />

must await high resolution sampling in future drilling.<br />

Never<strong>the</strong>less, <strong>the</strong> current data show that zone P1a is<br />

preserved throughout <strong>the</strong> Krishna-Godavari Basin during a<br />

period of local volcanic inactivity prior <strong>to</strong> eruptions of <strong>the</strong><br />

JOUR.GEOL.SOC.INDIA, VOL.78, NOV. 2011


DECCAN VOLCANISM LINKED TO THE CRETACEOUS-TERTIARY BOUNDARY MASS EXTINCTION 409<br />

Fig.10. ONGC well ELM-A with biostratigraphy, species occurrences, phase-2 and phase-3 mega-flows plotted against lithostratigraphy<br />

and e-logs (gamma and resistivity). Note <strong>the</strong>re are five thin phase-3 mega-flows, but only one in phase-2 below <strong>the</strong> KTB, which<br />

is likely due <strong>to</strong> erosion.<br />

last <strong>Deccan</strong> phase-3, as also observed in Rajahmundry<br />

quarries and Jhilmili, Madhya Pradesh (Keller et al. 2008,<br />

2009a, b).<br />

Volcanic phase-3 intertrappeans: Zone P1b<br />

Volcanic phase-3 began near <strong>the</strong> C29r/C29N boundary<br />

(Knight et al. 2003, 2005, Baksi, 2005; Chenet et al. 2007,<br />

2008), correlative with <strong>the</strong> zone P1a/P1b boundary. In <strong>the</strong><br />

Krishna-Godavari Basin, <strong>the</strong> three <strong>to</strong> four intertrappeans of<br />

phase-3 all contain planktic foraminiferal assemblages<br />

typical of zone P1b, which marks <strong>the</strong> interval between <strong>the</strong><br />

extinction of P. eugubina and P. longiapertura and <strong>the</strong> first<br />

appearance of Subbotina varianta (Figs. 7-13, Plates 4, 5).<br />

Faunal assemblages show increased abundances of G.<br />

daubjergensis, Guembelitria and biserial species and better<br />

preservation than Maastrichtian assemblages from<br />

intertrappeans of phase-2. However, <strong>the</strong>re are variations<br />

depending on <strong>the</strong> local depositional environment, such as<br />

in well MTP-A where few species are present due <strong>to</strong> poor<br />

preservation in sandy shale (Fig. 12).<br />

Danian above Phase-3 intertrappeans: Zone P1c or P2<br />

Sediments above volcanic phase-3 contain <strong>the</strong> first<br />

significantly more diverse Danian assemblages with <strong>the</strong> first<br />

larger (> 150 mm) morphotypes after <strong>the</strong> KTB mass<br />

extinction. In most K-G Basin wells, <strong>the</strong> assemblage is<br />

indicative of zone P1c with common subbotinids and <strong>the</strong><br />

first appearances of Praemurica inconstans and Subbotina<br />

JOUR.GEOL.SOC.INDIA, VOL.78, NOV. 2011


410 G. KELLER AND OTHERS<br />

Fig.11. ONGC well RZL-A with biostratigraphy, species occurrences, phase-2 and phase-3 mega-flows plotted against lithostratigraphy<br />

and e-logs (gamma and resistivity). Note <strong>the</strong> relatively thin mega-flows in phase-2, but absence of foraminifera in <strong>the</strong> intertrappean<br />

due <strong>to</strong> dissolution. Note also <strong>the</strong> unusually thick (55 m) mega-flow in phase-3, which likely is due <strong>to</strong> <strong>to</strong>pographic variation.<br />

varianta (Figs. 7, 9, 12). But in some wells <strong>the</strong>re is a<br />

significant hiatus with zone P2 overlying P1b, as indicated<br />

by <strong>the</strong> presence of assemblages with P. uncinata, M.<br />

angulata, M. praeangulata, A. strabocella, and S.<br />

triangularis (e.g., CTP-A, ELM-A, RZL-A, PNM-A, G-4-<br />

F (Figs. 1C, 10, 11, 13, 14). This hiatus may be related <strong>to</strong><br />

local <strong>to</strong>pographic variations, as suggested by <strong>the</strong> locations<br />

of <strong>the</strong> wells in shallower water (PNM-A), o<strong>the</strong>r hiatuses<br />

(G-4-F), and anomalously thick basalt mega-flows (RZL-A<br />

and CTP-A).<br />

BIOTIC EFFECTS OF DECCAN VOLCANISM<br />

Biotic effects of <strong>the</strong> world’s largest and longest lava flows<br />

Planktic foraminifera in intertrappeans record <strong>the</strong><br />

environmental conditions after <strong>the</strong> eruption of each megaflow.<br />

In <strong>the</strong> K-G Basin, this documentation is hampered by<br />

<strong>the</strong> limitation of samples and poor preservation. The former<br />

is largely a function of <strong>the</strong> difficulties <strong>to</strong> recover<br />

intertrappean sediments sandwiched between basalt flows<br />

in deep wells, and <strong>the</strong> latter is mainly a function of ocean<br />

acidification (acid rain) due <strong>to</strong> <strong>the</strong> volcanic eruptions.<br />

Because of <strong>the</strong> sporadic faunal records in individual wells,<br />

and because nine wells analyzed are within close proximity<br />

<strong>to</strong> each o<strong>the</strong>r in <strong>the</strong> Krishna-Godavari Basin (Fig. 1C), a<br />

composite was assembled based on nine wells with all data<br />

integrated in<strong>to</strong> <strong>the</strong> litholog of PLK-A (Fig. 15). All wells<br />

can be easily correlated based on biostratigraphy, gamma<br />

and resistivity logs, and <strong>the</strong> mega-flows below and above<br />

<strong>the</strong> KTB (Fig. 4). This dataset offers a glimpse in<strong>to</strong> <strong>the</strong><br />

environmental conditions and stresses associated with <strong>the</strong><br />

world’s largest and longest lava flows.<br />

JOUR.GEOL.SOC.INDIA, VOL.78, NOV. 2011


DECCAN VOLCANISM LINKED TO THE CRETACEOUS-TERTIARY BOUNDARY MASS EXTINCTION 411<br />

Fig.12. ONGC well MTP-A with biostratigraphy, species occurrences, phase-2 and phase-3 mega-flows plotted against lithostratigraphy<br />

and e-logs (gamma and resistivity). Core segment shown is from <strong>the</strong> intertrappean just below <strong>the</strong> mega-lava flow in phase-3. In<br />

this relatively shallow sandy environment, dissolution reduced overall species diversity and no species were recovered from<br />

phase-2 intertrappeans where dissolution is strongest.<br />

Onset of high-stress below phase-2 mega-flows<br />

During <strong>the</strong> late Maastrichtian, <strong>the</strong> Krishna-Godavari<br />

Basin wells show normal middle neritic diversity (28<br />

species) during zone CF2-CF3, except for a decrease in <strong>the</strong><br />

middle of this zone. Whe<strong>the</strong>r or not this diversity low is due<br />

<strong>to</strong> <strong>Deccan</strong> volcanism or poor preservation is uncertain,<br />

although it is likely due <strong>to</strong> culling of suspected down-core<br />

contaminants and poor preservation. Near <strong>the</strong> end of this<br />

interval, which is likely zone CF2, a 4 m long core in well<br />

NSP-A details <strong>the</strong> stress conditions immediately preceding<br />

phase-2 mega-flows (Fig.8). At <strong>the</strong> base of this core, diversity<br />

is normal at 28 species (typical for middle neritic depths,<br />

Keller and Abramovich, 2009), rapidly decreases <strong>to</strong> 18<br />

species by <strong>the</strong> middle of <strong>the</strong> core and drops <strong>to</strong> a low of 14<br />

species in <strong>the</strong> uppermost 2 m below <strong>the</strong> first mega-flow.<br />

This 50% faunal crash is likely due <strong>to</strong> environmental stresses<br />

attributable <strong>to</strong> increasing volcanism leading <strong>to</strong> <strong>the</strong> acme that<br />

corresponds <strong>to</strong> <strong>the</strong> mega-flows of phase-2 in <strong>the</strong> Krishna-<br />

Godavari Basin. A cosmic spherule was detected at 3365 m<br />

(Fig. 8). Such millimeter-sized metallic spherules are not<br />

uncommon in sediments of any age (Keller et al. 1983). They<br />

originate from extraterrestrial materials that melted during<br />

high-velocity entry in<strong>to</strong> <strong>the</strong> atmosphere. In contrast, glassy<br />

impact spherules (e.g., Chicxulub impact) result from hypervelocity<br />

impacts of large meteorites with earth’s surface.<br />

Maximum high-stress in intertrappeans of phase-2<br />

Intertrappean sediments of phase-2 (C29r, zones CF1)<br />

in <strong>the</strong> Krishna-Godavari Basin wells record severe<br />

environmental stresses in <strong>the</strong> aftermath of each of <strong>the</strong> four<br />

mega-flows, as earlier observed by Jaiprakash et al. (1993).<br />

Just 8 species were found after <strong>the</strong> first mega-flow, 13 and<br />

JOUR.GEOL.SOC.INDIA, VOL.78, NOV. 2011


412 G. KELLER AND OTHERS<br />

Fig.13. ONGC well PNM-A with biostratigraphy, species occurrences, and phase-3 mega-flows plotted against lithostratigraphy and e-<br />

logs (gamma and resistivity). Note that in this relatively shallow, sandy locality, <strong>the</strong> lower lava flows could not be identified<br />

because <strong>the</strong>y are very small and wea<strong>the</strong>red, or absent. Foraminifera are generally absent in <strong>the</strong> sandy layers.<br />

JOUR.GEOL.SOC.INDIA, VOL.78, NOV. 2011


DECCAN VOLCANISM LINKED TO THE CRETACEOUS-TERTIARY BOUNDARY MASS EXTINCTION 413<br />

Fig.14. ONGC offshore well G-4-F with biostratigraphy, species occurrences and phase 2 and phase-3 mega-flows plotted against<br />

lithostratigraphy and e-logs (gamma and resistivity). Note <strong>the</strong>re is a single phase-2 megaflow overlying lower Maastrichtian zone<br />

CF6 (69.08-69.61 Ma) faunal assemblages indicating that <strong>the</strong> upper Maastrichtian below <strong>the</strong> mega-flow is missing. Poor preservation<br />

and low diversity zone P1a-P1b assemblages are identified between phase-2 and phase-3. Ano<strong>the</strong>r major hiatus is above <strong>the</strong><br />

single phase-3 mega-flow.<br />

12 species after <strong>the</strong> second and third mega-flows (Fig. 15).<br />

However, <strong>the</strong>se data are not culled for down-core<br />

contamination and reworking. A culled record, eliminating<br />

all isolated single occurrences (open circles; Fig. 15) leaves<br />

just 7 <strong>to</strong> 8 species (black circles) in <strong>the</strong> first and second<br />

intertrappeans, and six species in <strong>the</strong> third intertrappean.<br />

This indicates a 50% reduction from <strong>the</strong> assemblages in <strong>the</strong><br />

infratrappean below, which already suffered a 50% faunal<br />

crash just prior <strong>to</strong> <strong>the</strong> onset of mega-flows in <strong>Deccan</strong><br />

phase-2 (Figs. 8, 15).<br />

We may argue that this is a crude estimate and that <strong>the</strong><br />

actual number of survivors may be higher or lower. However,<br />

since most of <strong>the</strong> species in this survivor group are among<br />

<strong>the</strong> environmentally most adaptable (Guembelitria cretacea,<br />

Heterohelix globulosa, Rugoglobigerina rugosa, Trinitella<br />

scotti, Globotruncana arca, G. aegyptiaca, G. dupeublei,<br />

G. conica), and two are KTB survivors (G. cretacea and H.<br />

globulosa), this estimate is probably not far off <strong>the</strong> mark. A<br />

more accurate assessment must await new coring of <strong>the</strong><br />

intertrappeans.<br />

Evidence of <strong>the</strong> detrimental effects of phase-2 volcanism<br />

has also been observed in <strong>the</strong> middle <strong>to</strong> inner neritic marine<br />

environment of <strong>the</strong> Um Sohryngkew River in Meghalaya,<br />

NE India (Gertsch et al. 2011). At this locality, about<br />

800 km from <strong>the</strong> main <strong>Deccan</strong> volcanic province, <strong>the</strong> last 4<br />

m of sediments below <strong>the</strong> KT boundary were deposited in<br />

nannofossil Micula prinsii zone (Garg et al. 2006), which is<br />

equivalent <strong>to</strong> C29R and zones CF1-CF2 below <strong>the</strong> KTB<br />

(Fig. 6). In this interval, <strong>the</strong> disaster opportunist<br />

Guembelitria cretacea dominates (~95%), with only rare<br />

and sporadic occurrences of 24 o<strong>the</strong>r <strong>Cretaceous</strong> species.<br />

But in <strong>the</strong> last meter below <strong>the</strong> KTB, all but seven species<br />

JOUR.GEOL.SOC.INDIA, VOL.78, NOV. 2011


414 G. KELLER AND OTHERS<br />

disappeared. This pattern is very similar <strong>to</strong> that observed in<br />

<strong>the</strong> Krishna-Godavari Basin and suggests <strong>the</strong> same<br />

detrimental environmental effects of phase-2 volcanism at<br />

a large distance from <strong>the</strong> volcanic province.<br />

Based on <strong>the</strong> current Krishna-Godavari Basin dataset<br />

and <strong>the</strong> similar record observed in <strong>the</strong> Meghalaya section,<br />

we conclude that in India and its surrounding area, <strong>the</strong><br />

<strong>Cretaceous</strong> planktic foraminifera were already near<br />

extinction before <strong>the</strong> KT boundary as a result of <strong>the</strong> pulsed<br />

volcanic eruptions of phase-2 that created <strong>the</strong> world’s largest<br />

and longest lava flows (Fig. 15). This data also indicates<br />

that high-stress environmental conditions continued<br />

unabated between <strong>the</strong> pulsed mega-eruptions and prevented<br />

marine recovery. Such prolonged high-stress conditions<br />

likely resulted from <strong>the</strong> rapid succession of <strong>Deccan</strong> eruptions<br />

in phase-2 and <strong>the</strong> corollary effects of gas emissions, climate<br />

change and increased wea<strong>the</strong>ring rates (Gertsch et al. 2011).<br />

KT boundary Event<br />

In all eleven Krishna-Godavari Basin wells analyzed <strong>the</strong><br />

intertrappean above <strong>the</strong> phase-2 mega-flows contain <strong>the</strong><br />

earliest Danian zone P1a assemblage and occasionally rare<br />

reworked <strong>Cretaceous</strong> species (Figs. 7-15) consistent with<br />

previous observations in Rajahmundry quarries (Keller et<br />

al. 2008; Malarkodi et al. 2010). This demonstrates that <strong>the</strong><br />

mass extinction of <strong>Cretaceous</strong> species occurred prior <strong>to</strong><br />

deposition of <strong>the</strong> intertrappean and was coeval with <strong>Deccan</strong><br />

volcanic phase-2, suggesting a cause-and-effect relationship.<br />

High-resolution core sampling will be necessary <strong>to</strong> evaluate<br />

<strong>the</strong> presence of <strong>the</strong> boundary clay (zone P0) and lower part<br />

of zone P1a, measure d 13 C values and evaluate <strong>the</strong> presence<br />

of Ir and o<strong>the</strong>r PGEs.<br />

Continued high-stress in early Danian zone P1a intertrappean<br />

fauna<br />

Intertrappeans between <strong>Deccan</strong> phase-2 and phase-3<br />

mega-flows were deposited in C29R above <strong>the</strong> KTB, which<br />

is equivalent <strong>to</strong> planktic foraminiferal zone P1a, or about<br />

380 ky of <strong>the</strong> basal Danian (Fig. 6), as also indicated by<br />

40 Ar/ 39 Ar dating of phase-3 mega-flows (Knight et al. 2003,<br />

2005; Baksi, 2005). Evolution of Danian species in this<br />

intertrappean follows <strong>the</strong> same pattern as observed globally,<br />

Fig.15. Composite species ranges of nine wells in <strong>the</strong> Krishna-Godavari Basin plotted against biostratigraphy and phase-2 and phase-3<br />

lava flows of <strong>the</strong> PLK-A well. Cored intervals at <strong>the</strong> base of <strong>the</strong> section, below <strong>the</strong> first phase-2 mega-flow and below <strong>the</strong> last<br />

mega-flow of phase-3 record <strong>the</strong> most reliable species richness data. A maximum of 28 species in <strong>the</strong> upper Maastrichtian is<br />

typical for middle neritic environments (~100 m depth). Note <strong>the</strong> mass extinction began with <strong>the</strong> onset of phase-2 volcanism<br />

(50% drop in species richness), with ano<strong>the</strong>r 50% drop after <strong>the</strong> first lava flow, and was complete by <strong>the</strong> last mega-flow at or near<br />

<strong>the</strong> KTB. Phase-2 intertrappeans: open circles = contamination and reworking, black circles = in situ species.<br />

JOUR.GEOL.SOC.INDIA, VOL.78, NOV. 2011


DECCAN VOLCANISM LINKED TO THE CRETACEOUS-TERTIARY BOUNDARY MASS EXTINCTION 415<br />

Fig.16. Depth ranking of species in high diversity optimum planktic foraminiferal assemblages from outer neritic <strong>to</strong> open oceanic<br />

environments of <strong>the</strong> late Maastrichtian. Most large specialized (k-strategy) species evolved and thrived in <strong>the</strong>rmocline and<br />

sub<strong>the</strong>rmocline deep depths. Smaller (r-strategy) species, <strong>to</strong>lerant of fluctuations in oxygen, salinity and temperature, thrived in<br />

surface, and <strong>the</strong>rmocline depths. (modified from Keller and Abramovich, 2009). 1. Pseudoguembelina palpebra, 2. Heterohelix<br />

planta, 3. Pseudoguembelina hariaensis, 4. Heterohelix navarroensis, 5. Pseudoguembelina costulata, 6. Pseudoguembelina<br />

kempensis, 7. Pseudoguembelina excolata, 8. Rugoglobigerina rotundata, 9. Rugoglobigerina rugosa, 10. Pseudotextularia<br />

elegans, 11. Racemiguembelina fructicosa, 12. Pseudotextularia deformis, 13. Heterohelix globulosa, 14. Plummertita<br />

hantkeninoides 15. Globotruncana aegyptiaca, 16. Rugoglobigerina scotti, 17. Planoglobulina acervulinoides, 18. Hedbergella<br />

monmou<strong>the</strong>nsis, 19. Globigerinelloides aspera, 20. Heterohelix labellosa, 21. Globotruncana arca, 22. Contusotruncana<br />

contusa, 23. Globotruncanita stuarti, 24. Globotruncanita stuartiformis, 25. Heterohelix rajagopalani*, 26. Gublerina acuta,<br />

27. Globotruncanella citae, 28. Laeviheterohelix glabrans, 29. Abathomphalus mayaroensis, 30. Gublerina cuvillieri,<br />

31. Planoglobulina multicamerata.* rare or not present in neritic environments.<br />

although first appearances of some species may differ due<br />

<strong>to</strong> poor preservation or large sample spacing (Figs. 7-14).<br />

All evolving early Danian species are very small (< 100<br />

mm and frequently < 63 mm), with simple chamber<br />

arrangements and unornamented morphologies that reflect<br />

adaption for survival in highly stressed environments (e.g.,<br />

Keller and Pardo, 2004; Pardo and Keller, 2008). The fact<br />

that <strong>the</strong>se assemblages mirror <strong>the</strong> global evolutionary pattern<br />

demonstrates that environmental conditions after <strong>the</strong> main<br />

phase-2 of <strong>Deccan</strong> volcanism remained stressed in India<br />

and globally.<br />

Continued high-stress in early Danian (zone P1b) <strong>Deccan</strong><br />

Phase-3<br />

Faunal assemblages in phase-3 intertrappeans are similar<br />

<strong>to</strong> zone P1a, except for <strong>the</strong> disappearance of P. eugubina,<br />

<strong>the</strong> decreased abundance of early zone P1a species (e.g.,<br />

Parvularugoglobigerina extensa, Eoglobigerina edita,<br />

Woodringina clay<strong>to</strong>nensis, W. horners<strong>to</strong>wnensis, E.<br />

eobulloides), and increased abundance of o<strong>the</strong>r species<br />

(Figs. 8-14). Species sizes remained small, with<br />

morphologies generally < 150 mm, as also observed globally<br />

(Keller, 1988; 1989; Luciani, 2002; Coccioni and Luciani,<br />

JOUR.GEOL.SOC.INDIA, VOL.78, NOV. 2011


416 G. KELLER AND OTHERS<br />

Fig.17. Depth ranking of species in high-stress, low diversity planktic foraminiferal assemblages from middle <strong>to</strong> inner neritic environments<br />

of <strong>the</strong> late Maastrichtian. High biotic stress selectively eliminates large specialized (k-strategy) species from subsurface and<br />

<strong>the</strong>rmocline depths, leaving impoverished assemblages. Smaller (r-strategy) species thrive, particularly <strong>the</strong> low oxygen <strong>to</strong>lerant<br />

heterohelicids. These biotic effects are also observed in high-stress conditions related <strong>to</strong> major volcanism. (Modified from<br />

Keller and Abramovich, 2009). 1. Heterohelix planta, 2. Pseudoguembelina hariaensis, 3. Guembelitria cretacea, 4. Heterohelix<br />

navarroensis, 5. Pseudoguembelina costulata, 6. Pseudotextularia elegans, 7. Rugoglobigerina rugosa, 8. Heterohelix globulosa,<br />

9. Globigerinelloides aspera, 10. Hedbergella monmou<strong>the</strong>nsis, 11. Contusotruncana contusa, 12. Globotruncana arca,<br />

13. Globotruncana aegyptiaca, 14. Abathomphalus mayaroensis.<br />

2006; Keller et al. 2007a, b, 2009e). No major differences<br />

are observed between faunal assemblages of <strong>the</strong> three<br />

intertrappeans, which suggests that environmental conditions<br />

remained <strong>to</strong>lerable during volcanic phase-3. Alternatively,<br />

since <strong>the</strong>se early Danian species evolved during high-stress<br />

conditions, <strong>the</strong>y were primed for survival, unlike most<br />

species in <strong>the</strong> late Maastrichtian, which were highly<br />

specialized and adapted for narrow ecological niches<br />

(Abramovich et al. 2003; Keller and Abramovich, 2009).<br />

Recovery after phase-3 <strong>Deccan</strong> volcanism<br />

Planktic foraminiferal assemblages from sediments<br />

above <strong>the</strong> last phase-3 mega-flows reveal generally larger<br />

species sizes and increased diversity. On a global basis, <strong>the</strong><br />

first Danian species with morphology sizes > 150 mm are<br />

generally observed near <strong>the</strong> end of zone P1b and full marine<br />

recovery did not occur until zone P1c (Keller, 1988, 1989;<br />

Keller et al. 2009e). It is tempting <strong>to</strong> speculate that <strong>the</strong> long<br />

delay in <strong>the</strong> global ecosystem recovery was due <strong>to</strong> <strong>Deccan</strong><br />

volcanism. Studies are now underway <strong>to</strong> investigate this<br />

possibility.<br />

DISCUSSION<br />

Planktic Foraminifera – Proxies for Environmental Change<br />

We can assess <strong>the</strong> biologic effects of <strong>Deccan</strong> volcanism<br />

based on planktic foraminiferal assemblages, which are<br />

highly sensitive <strong>to</strong> environmental changes. At <strong>the</strong> KTB<br />

planktic foraminifera suffered <strong>the</strong> most devastating mass<br />

extinction with just a few environmentally more <strong>to</strong>lerant<br />

JOUR.GEOL.SOC.INDIA, VOL.78, NOV. 2011


DECCAN VOLCANISM LINKED TO THE CRETACEOUS-TERTIARY BOUNDARY MASS EXTINCTION 417<br />

species surviving for a short time in<strong>to</strong> <strong>the</strong> early Danian and<br />

<strong>the</strong> disaster opportunist Guembelitria cretacea <strong>the</strong> sole longterm<br />

survivor. Consequently, this microfossil group has<br />

become <strong>the</strong> strongest paleon<strong>to</strong>logical proxy for evaluating<br />

<strong>the</strong> biological effects of catastrophes, whe<strong>the</strong>r meteorite<br />

impacts, volcanism or climate change (e.g., Pardo and Keller,<br />

2008; Coccioni and Luciani, 2006; Keller and Abramovich,<br />

2009; Abramovich et al. 2003, 2010).<br />

Planktic foraminifera can be grouped in<strong>to</strong> surface,<br />

intermediate (subsurface) and deep dwellers based on stable<br />

iso<strong>to</strong>pic ranking (Keller, 2001; Abramovich et al. 2003;<br />

Keller and Abramovich, 2009). Optimum assemblages are<br />

characterized by low diversity surface dwellers, high<br />

diversity intermediate dwellers and very low diversity deep<br />

dwellers (below <strong>the</strong>rmocline) (Fig. 16). The intermediate<br />

assemblages consist of many large, ornamented and<br />

highly specialized taxa (K-strategists), adapted <strong>to</strong> specific<br />

ecological niches and are <strong>the</strong>refore most strongly<br />

affected by environmental changes, such as variations in<br />

temperature, salinity, oxygen, nutrients and acidity of<br />

<strong>the</strong> water column. Any variations in <strong>the</strong>se fac<strong>to</strong>rs can<br />

result in high-stress conditions whe<strong>the</strong>r in open oceans,<br />

restricted basins, marginal marine settings or volcanic<br />

activity (Keller and Abramovich, 2009; Kidder and Worsley,<br />

2010).<br />

During high-stress conditions diversity drops most<br />

strongly among intermediate dwellers (K-strategists) leaving<br />

a survivor assemblage of dwarfed species (Lilliput effect)<br />

and less specialized (r-strategist) taxa (e.g., heterohelicids,<br />

guembelitrids, globigerinellids, and few rugoglobigerinids,<br />

pseudoguembelinids, globotruncanids, such as R. rugosa,<br />

P. costulata, G. arca, Fig. 17; Keller and Abramovich, 2009).<br />

Increasing biotic stress results in <strong>the</strong> complete disappearance<br />

K-strategists, <strong>the</strong> dwarfing of species more <strong>to</strong>lerant of<br />

environmental changes (r-strategists) and dominance by low<br />

oxygen <strong>to</strong>lerant small heterohelicids. At <strong>the</strong> extreme end of<br />

<strong>the</strong> biotic response are volcanically influenced environments<br />

which cause <strong>the</strong> same detrimental biotic effects as observed<br />

in <strong>the</strong> aftermath of <strong>the</strong> KTB mass extinction, including <strong>the</strong><br />

disappearance of most species and blooms of <strong>the</strong> disaster<br />

opportunist Guembelitria (Fig. 18, Abramovich and Keller,<br />

2002; Abramovich et al. 2002; Keller, 2003; Keller et al.<br />

2007a). Volcanically induced high-stress conditions have<br />

been documented from Ninetyeast Ridge, Andean volcanism<br />

and <strong>the</strong> <strong>Deccan</strong> volcanic province (Keller, 2003; Keller et<br />

al. 2007a; Gertsch et al. 2011; this study).<br />

We can interpret <strong>the</strong> biologic response of planktic<br />

foraminifera <strong>to</strong> <strong>Deccan</strong> volcanism based on global<br />

observations of depth-ranked species and stress conditions<br />

in various habitats (Figs. 16-18). In <strong>the</strong> K-G Basin <strong>the</strong> change<br />

from optimum <strong>to</strong> high-stress assemblage occurred in <strong>the</strong> 4<br />

m below <strong>the</strong> <strong>Deccan</strong> phase-2 mega-flows (Fig. 8). The 50%<br />

gradual decrease in species richness indicates increasing<br />

stress, probably related <strong>to</strong> increasing volcanic activity<br />

leading up <strong>to</strong> Earth’s largest and longest mega-flows.<br />

During phase-2 mega-flows super-stress conditions<br />

approached <strong>the</strong> catastrophic with ano<strong>the</strong>r 50% drop in<br />

species richness leaving just a few generally dwarfed<br />

survivors in intertrappeans and <strong>the</strong> mass extinction was<br />

complete at <strong>the</strong> last mega-flow of phase-2 (Fig. 15). In<br />

general, Guembelitria and Heterohelix species dominate<br />

super-stress <strong>to</strong> catastrophic conditions (Fig. 18). Their rarity<br />

and absence in <strong>the</strong> Krishna-Godavari Basin wells is an<br />

artifact of preservation. Above <strong>the</strong> KTB <strong>the</strong> low diversity<br />

early Danian assemblages of small, unornamented species<br />

with simple chamber arrangements persisted until after<br />

<strong>Deccan</strong> volcanism ended with phase-3. This long-delayed<br />

recovery in <strong>the</strong> marine ecosystem has long been an<br />

enigma. The current K-G Basin data suggest that <strong>the</strong><br />

delayed recovery may have been due <strong>to</strong> continued <strong>Deccan</strong><br />

volcanism.<br />

Climate Change and <strong>Deccan</strong> <strong>Volcanism</strong><br />

Major global climate changes occurred during <strong>the</strong> late<br />

Maastrichtian C29R (zones CF1-CF2), including maximum<br />

<strong>Cretaceous</strong> cooling at <strong>the</strong> end of zone CF3, rapid global<br />

warming beginning in zone CF2 reaching a maximum in<br />

zone CF1, followed by rapid cooling in <strong>the</strong> upper part of<br />

CF1 and across <strong>the</strong> KTB (e.g., Li and Keller, 1998c; Kucera<br />

and Malmgren, 1998; Olsson et al. 2001; Abramovich and<br />

Keller, 2003; Wilf et al. 2003; Nordt et al. 2003; Keller and<br />

Abramovich, 2009; Kidder and Worsley, 2010). This global<br />

climate change documented in South Atlantic DSDP Site<br />

525 is representative for India, which was at <strong>the</strong> equivalent<br />

sou<strong>the</strong>rn latitude (Fig. 19). During <strong>the</strong> global warming in<br />

CF2 <strong>to</strong> CF1, deeper water temperatures recorded in benthic<br />

foraminifera increased by about 4°C, whereas surface water<br />

temperatures fluctuated between 15 and 17°C. At <strong>the</strong> same<br />

time planktic foraminifera suffered under high-stress<br />

conditions leading <strong>to</strong> species dwarfing (60% of fauna),<br />

decreased diversity and abundance of specialized large<br />

species (e.g., intermediate dwellers). Global temperatures<br />

rapidly cooled about 50 ky or 100 ky prior <strong>to</strong> <strong>the</strong> KTB (time<br />

scale of Gradstein et al. 2004). Cooling was accompanied<br />

by a rapid decline in surface, intermediate and deep dwellers<br />

and ended in <strong>the</strong> KTB mass extinction (Keller and<br />

Abramovich, 2009; Keller et al. 2009e). This end-<br />

Maastrichtian global cooling and mass extinction appears<br />

coeval and likely related <strong>to</strong> <strong>the</strong> main <strong>Deccan</strong> phase-2 in <strong>the</strong><br />

Krishna-Godavari Basin.<br />

JOUR.GEOL.SOC.INDIA, VOL.78, NOV. 2011


418 G. KELLER AND OTHERS<br />

Fig.18. The effects of increasing environmental stress upon planktic foraminiferal assemblages from optimum <strong>to</strong> catastrophe conditions<br />

shows <strong>the</strong> successive elimination of large, specialized k-strategy species, <strong>the</strong> survival of small r-strategy species, <strong>the</strong> overall<br />

dwarfing of <strong>the</strong>se species and <strong>the</strong>ir great abundance. All of <strong>the</strong>se fac<strong>to</strong>rs characterize <strong>the</strong> Lilliput effect and are characteristic also<br />

of high-stress conditions associated with major volcanism (modified from Keller and Abramovich, 2009).<br />

Environmental Effects of <strong>Deccan</strong> <strong>Volcanism</strong><br />

Environmental consequences of <strong>Deccan</strong> phase-2<br />

eruptions were likely devastating. Based on rare gas bubbles<br />

preserved in <strong>Deccan</strong> volcanic rocks, Self et al. (2008b)<br />

estimate annual gas rates released at many times <strong>the</strong> rate of<br />

anthropogenic emissions of SO 2<br />

and more than an order of<br />

magnitude greater than <strong>the</strong> current global background<br />

volcanic emission rate. Chenet et al. (2007, 2008) estimated<br />

gas emissions based on <strong>the</strong> volume of <strong>the</strong> largest 30 <strong>Deccan</strong><br />

eruption pulses, with each pulse injecting 30-150 GT of SO 2<br />

gas over a very short time (decades). Thus each of <strong>the</strong> 30<br />

<strong>Deccan</strong> eruption pulses could have injected quantities of<br />

SO 2<br />

equivalent <strong>to</strong> that of <strong>the</strong> Chicxulub impact (e.g. 50-500<br />

GT). Indeed, <strong>the</strong> main phase-2 <strong>Deccan</strong> eruptions are<br />

estimated <strong>to</strong> have released 30 <strong>to</strong> 100 times <strong>the</strong> amount of<br />

SO 2<br />

released by <strong>the</strong> Chicxulub impact. Given <strong>the</strong>se<br />

estimates, <strong>the</strong> environmental effects of <strong>Deccan</strong> volcanism<br />

were likely orders of magnitude worse than those of <strong>the</strong><br />

Chicxulub impact.<br />

CO 2<br />

and SO 2<br />

gas emissions<br />

The global greenhouse warming of about 3-4°C<br />

beginning in zone CF2 and ending in zone CF1 resulted in<br />

major biologic stress for <strong>the</strong> Maastrichtian fauna leading<br />

<strong>to</strong> decreased species population abundances, decreased<br />

diversity, and species dwarfing (Fig. 19). Researchers<br />

have attributed this warm event <strong>to</strong> <strong>Deccan</strong> volcanism.<br />

However, climate models predict at most a 2°C warming,<br />

suggesting that <strong>the</strong> CO 2<br />

emissions from <strong>Deccan</strong> volcanism<br />

alone would have been insufficient <strong>to</strong> cause this warming<br />

(DeCon<strong>to</strong> et al. 2000; Donnadieu et al. 2006). In <strong>the</strong> absence<br />

of any o<strong>the</strong>r likely CO 2<br />

source for <strong>the</strong> rapid greenhouse<br />

warming <strong>Deccan</strong> volcanism remains <strong>the</strong> only realistic<br />

event.<br />

Global cooling followed <strong>the</strong> warm event during <strong>the</strong> last<br />

~50-100 ky of <strong>the</strong> Maastrichtian and may be correlative with<br />

<strong>the</strong> mega-flows of phase-2 <strong>Deccan</strong> volcanism (Fig. 19). This<br />

cooling could have been <strong>the</strong> result of SO 2<br />

gas released by<br />

<strong>Deccan</strong> volcanism. SO 2<br />

injected in<strong>to</strong> <strong>the</strong> stra<strong>to</strong>sphere forms<br />

sulfate aerosol particulates, which act <strong>to</strong> reflect incoming<br />

solar radiation and cause global cooling. Because sulfate<br />

aerosol has a short lifespan in <strong>the</strong> atmosphere, <strong>the</strong> cooling<br />

would be short-term (years <strong>to</strong> decades), unless repeated<br />

injections from volcanic eruptions replenished atmospheric<br />

JOUR.GEOL.SOC.INDIA, VOL.78, NOV. 2011


DECCAN VOLCANISM LINKED TO THE CRETACEOUS-TERTIARY BOUNDARY MASS EXTINCTION 419<br />

Fig.19. Biotic response <strong>to</strong> greenhouse warming at South Atlantic DSDP Site 525 (data from Li and Keller, 1998c; Abramovich and<br />

Keller, 2003). During <strong>the</strong> greenhouse warming, diversity (H’) dropped, large specialized (k-strategy) species reached adulthood<br />

at less than 50% of <strong>the</strong>ir former adult size with up <strong>to</strong> 60% of <strong>the</strong> specimens dwarfed, and low oxygen <strong>to</strong>lerant Heterohelix species<br />

temporarily decreased in abundance. Dwarfed species decreased at <strong>the</strong> end of <strong>the</strong> Maastrichtian although no recovery occurred<br />

and k-strategy species remained rare (


420 G. KELLER AND OTHERS<br />

of <strong>Deccan</strong> volcanism: <strong>the</strong> main phase-2 at <strong>the</strong> end of <strong>the</strong><br />

Maastrichtian accounts for about 80% and phase-3 in <strong>the</strong><br />

early Danian for about 14% of <strong>the</strong> <strong>to</strong>tal <strong>Deccan</strong> volume.<br />

<strong>Deccan</strong> main phase-2 volcanism and its corollary effects<br />

(warming, cooling, acid rain, ocean acidification, carbon<br />

crisis) are viable and <strong>the</strong> likely main cause of <strong>the</strong> KTB mass<br />

extinction. The shear volume and rapid rate of volcanic<br />

eruptions and <strong>the</strong> number of mega-flows make phase-2<br />

volcanism far more destructive than a single large impact.<br />

Instead of a single instantaneous catastrophe, <strong>the</strong> KTB<br />

killer was likely a cascade of rapid and massive volcanic<br />

eruptions that formed <strong>the</strong> largest and longest (1500 km) lava<br />

flows known on Earth, with SO 2<br />

and CO 2<br />

gas emissions<br />

estimated <strong>to</strong> have exceeded those of <strong>the</strong> Chicxulub impact<br />

by at least 30 times (Chenet et al. 2007, 2008). The last<br />

phase-3 of <strong>Deccan</strong> eruptions in <strong>the</strong> early Danian was much<br />

smaller (14% of <strong>to</strong>tal <strong>Deccan</strong> Traps) and caused no<br />

significant species extinctions, but resulted in high-stress<br />

conditions that may be <strong>the</strong> cause for <strong>the</strong> long delayed full<br />

marine recovery after <strong>the</strong> mass extinction. Major findings<br />

from <strong>the</strong> Krishna-Godavari Basin study support this<br />

scenario.<br />

<strong>Deccan</strong> phase-2 volcanism (~80% of <strong>to</strong>tal <strong>Deccan</strong> Traps)<br />

began in C29R and ended at <strong>the</strong> KTB. This interval of<br />

C29R is equivalent <strong>to</strong> zones CF1-CF2, which span <strong>the</strong><br />

last 280 ky of <strong>the</strong> Maastrichtian and is equivalent <strong>to</strong> most<br />

of Micula prinsii zone (Fig. 6). However, phase-2<br />

eruptions may have occurred over a much shorter time<br />

interval and mainly concentrated in zone CF1. <strong>Deccan</strong><br />

phase-2 ended with massive eruptions that formed<br />

four of Earth’s largest and longest lava mega-flows<br />

(~1500 km) across India and out in<strong>to</strong> <strong>the</strong> Bay of Bengal.<br />

· The KTB mass extinction began in <strong>Deccan</strong> phase-2 with<br />

a 50% faunal crash prior <strong>to</strong> <strong>the</strong> first of four mega-flows,<br />

followed by ano<strong>the</strong>r 50% crash after <strong>the</strong> first mega-flow,<br />

leaving just 6 <strong>to</strong> 8 survivors. No recovery occurred<br />

between mega-flows and <strong>the</strong> mass extinction was<br />

complete with <strong>the</strong> last mega-flows that marks <strong>the</strong> end of<br />

phase-2 and <strong>the</strong> KTB.<br />

The kill effect was likely due <strong>to</strong> <strong>the</strong> rapid, pulsed<br />

eruptions of massive CO 2<br />

and SO 2<br />

gas emissions, leading<br />

<strong>to</strong> high continental wea<strong>the</strong>ring rates, global warming,<br />

cooling, acid rains, ocean acidification and a carbon<br />

crisis that prevented recovery.<br />

After <strong>the</strong> KTB mass extinction, early Danian (zone P1a)<br />

assemblages in India mirror <strong>the</strong> global evolutionary<br />

pattern and global high-stress conditions.<br />

<strong>Deccan</strong> phase-3 volcanism (~14% of <strong>to</strong>tal <strong>Deccan</strong><br />

Traps) began in <strong>the</strong> early Danian near <strong>the</strong> C29R/C29N<br />

boundary, which is equivalent <strong>to</strong> <strong>the</strong> P1a/P1b boundary<br />

and also formed three <strong>to</strong> four of Earth’s largest and<br />

longest lava flows. The onset of phase-3 coincided with<br />

<strong>the</strong> extinctions of P. eugubina and P. longiapertura.<br />

<strong>Deccan</strong> phase-3 (zone P1b) reveals no major differences<br />

between faunal assemblages in <strong>the</strong> three intertrappeans<br />

or <strong>the</strong> global record. This suggests that environmental<br />

conditions remained <strong>to</strong>lerable, possibly because volcanic<br />

eruptions were less intense or separated by longer time<br />

intervals, and/or that early Danian species were well<br />

adapted <strong>to</strong> <strong>the</strong> environmental stresses.<br />

Recovery of <strong>the</strong> marine ecosystem occurred after <strong>the</strong><br />

last <strong>Deccan</strong> phase-3, as indicated by larger species<br />

morphotypes and evolution of more diverse assemblages.<br />

The long delayed post-KTB recovery in <strong>the</strong> marine<br />

environment has long been an enigma. Data from <strong>the</strong><br />

Krishna-Godavari Basin wells suggest that continued<br />

high-stress conditions caused by <strong>Deccan</strong> volcanism and<br />

its corollary effects may have prevented full recovery<br />

for over 0.5 Ma.<br />

Acknowledgments: This project would not have been<br />

possible without <strong>the</strong> support of former Direc<strong>to</strong>r<br />

(Exploration) Dr. D.K. Pande and <strong>the</strong> current Direc<strong>to</strong>r Dr.<br />

S.V. Rao of <strong>the</strong> Oil and Natural Gas Corporation Ltd., India.<br />

The senior author is deeply grateful for <strong>the</strong> permission <strong>to</strong><br />

study <strong>the</strong> Krishna-Godavari Basin wells that made this study<br />

possible. A special thanks <strong>to</strong> Dr. D.S.N. Raju for sharing his<br />

extensive knowledge of <strong>the</strong> Krishna-Godavari Basin wells,<br />

and <strong>to</strong> Dr. Sunil Bajpai who facilitated this study in many<br />

ways. We are also very grateful <strong>to</strong> Mr. D.K. Bharktya, DGM<br />

(Geology) RGl, Chennai, for taking SEMs of <strong>the</strong> foraminifera.<br />

This study is based upon work supported by <strong>the</strong> US National<br />

Science Foundation through <strong>the</strong> Continental Dynamics<br />

Program, Sedimentary Geology and Paleobiology Program<br />

and Office of International Science & Engineering’s India<br />

Program under NSF Grants EAR-0207407, EAR-0447171,<br />

and EAR-1026271.<br />

JOUR.GEOL.SOC.INDIA, VOL.78, NOV. 2011


DECCAN VOLCANISM LINKED TO THE CRETACEOUS-TERTIARY BOUNDARY MASS EXTINCTION 421<br />

Plate 1. SEM illustrations of late Maastrichtian, scale bar = 100 µm. 1. Globotruncanita stuarti (de Lapparent, 1918), Well RZL-A,<br />

3650 m. 2-4. Globotruncanita stuartiformis (Dalbiez, 1955), Wells ELM-A, 3590 m, NSP-A, 3363 m. 5-8. Globotruncana arca (Cushman,<br />

1926), Wells PNM-A, 2555 m, CTP-A, 3980 m, RZL-A, 3650 m. 9-12. GLobotruncana insignis Gandolfi, 1955, Wells PNM-A,<br />

2545 m, RZL-A, 3650 m. 13-16. Globotruncana rosetta (Carsey, 1926), Wells PNM-A, 2565 m, CTP-A, 4010 m, RZL-A, 3650 m.<br />

JOUR.GEOL.SOC.INDIA, VOL.78, NOV. 2011


422 G. KELLER AND OTHERS<br />

Plate 2. SEM illustrations of late Maastrichtian, scale bar = 100 µm. 1-2. Abathomphalus mayaroensis (Bolli, 1951)Well CTP-A,<br />

4010 m. 3-4. Rosita patelliformis (Gandolfi, 1955), Well CTP-A, 3960 m. 5-6. Globotruncanita conica (White, 1928), Wells NSP-A,<br />

3363 m, RZL-A, 3650 m. 7-8. Rosita walfishensis (Todd, 1970), Well CTP-A, 4010 m. 9-10. Globotruncan dupeublei (Caron et<br />

al.1984), Well RZL-A, 3650 m. 11-12. Glo<strong>to</strong>bruncanita pettersi (Gandolfi, 1955), Well PNM-A, 2565 m. 13-15. Rosita contusa (Cushman,<br />

1926), Well PNM-A, 2565 m.<br />

JOUR.GEOL.SOC.INDIA, VOL.78, NOV. 2011


DECCAN VOLCANISM LINKED TO THE CRETACEOUS-TERTIARY BOUNDARY MASS EXTINCTION 423<br />

Plate 3. SEM illustrations of late Maastrichtian, scale bar = 100 µm. 1. Racemiguembelina fructicosa (Egger, 1899), Well RZL-A, 3650<br />

m. 2. Racemiguembelina powelli Smith and Pessagno, 1973, Well MTP-A, 3160 m. 3-4. Planoglobulina brazoensis (Plummer, 1931),<br />

Well ELM-A, 3590 m. 5-6. Heterohelix labellosa, Nederbragt, 1990, Well NSP-A, 3339.2 m. 7. Pseudoguembelina palpebra Brönnimann<br />

and Brown, 1953, Well NSP-A, 3339.5 m. 8. Planoglobulina carseyae (Plummer, 1931), Well NSP-A, 3339.5 m. 9-10. Heterohelix<br />

globulosa (Ehrenberg), Well NSP-A, 3339.5 m, 3339.2 m. 11-12. Pseudoguembelina hariaensis Nederbragt, 1990, Well NSP-A,<br />

3339.5 m. 13. Heterohelix planata (Cushman, 1936), Well RZL-A, 3650 m. 14-15. Pseudoguembelina costulata (Cushman, 1938),<br />

Well NSP-A, 3339.5 m. 16. Pseudotextularia nuttalli (Rzehak, 1886), Well PNM-A, 2545 m. 17. Pseudotextularia elegans (Rzehak,<br />

1886), Well PNM-A, 2535 m.<br />

JOUR.GEOL.SOC.INDIA, VOL.78, NOV. 2011


424 G. KELLER AND OTHERS<br />

Plate 4. SEM illustrations of early Danian, scale bar = 100 µm. 1-7. Parasubbotina pseudobulloides (Plummer, 1926), Wells PNM-A,<br />

2345 m, Razole-1, 3385 m. 8-11. Subbotina triloculinoides (Plummer, 1926), Wells PNM-A, 2345 m, 2350 m, Razole-1, 3330 m.<br />

12-15. Subbotina varianta (Subbotina, 1953), Well PNM-A, 2250 m. 16. Globigerina (Eoglobigerina) tetragona (Morozova, 1961),<br />

Well PNM-A, 2295 m.<br />

JOUR.GEOL.SOC.INDIA, VOL.78, NOV. 2011


DECCAN VOLCANISM LINKED TO THE CRETACEOUS-TERTIARY BOUNDARY MASS EXTINCTION 425<br />

Plate 5. SEM illustrations of early Danian, scale bar = 100 µm. 1-2. Subbotina trivialis (Subbotina, 1953), Wells PNM-A, 2295 m,<br />

RZL-A, 3365 m. 3. Globoconusa daugjergensis (Brönnimann, 1953), Well PNM-A, 2295 m. 4-6. Eoglobigerina edita (Subbotina,<br />

1953), Well RZL-A, 3370 m. 7-9. Globigerina (Eoglobigerina) pentagona (Morozova, 1961), Well RZL-A, 3370 m. 10-11. Praemurica<br />

taurica (Morozova, 1961), Well PNM-A, 2345 m. 12, 16. Praemurica inconstans (Subbotina, 1953), Well PNM-A, 2350 m.<br />

13-15. Praemurica uncinata (Bolli, 1957), Wells PLK-A, 2565 m, PNM-A, 2190 m.<br />

JOUR.GEOL.SOC.INDIA, VOL.78, NOV. 2011


426 G. KELLER AND OTHERS<br />

References<br />

ABRAMOVICH, S., YOVEL-COREM, S., ALMOGI-LABIN, A. and<br />

BENJAMINI, C. (2010) Global climate change and planktic<br />

formaminiferal response in <strong>the</strong> Maastrichtian. Paleoceanography,<br />

v.25, PA2201, doi: 10.1029/2009PA001843.<br />

ABRAMOVICH, S. and KELLER, G. (2002) High stress late<br />

Maastrichtian Paleoenvironment in Tunisia: Inference from<br />

planktic foraminifera. Paleogeog., Paleoclimat., Paleoeco.,<br />

v.178, pp.145-164.<br />

ABRAMOVICH, S. and KELLER, G. (2003) Planktic foraminiferal<br />

response <strong>to</strong> latest Maastrichtian abrupt warm event a case study<br />

from mid-latitude DSDP Site 525. Mar. Micropal., v.48,<br />

pp.225-249.<br />

ABRAMOVICH, S., KELLER, G., ADATTE, T., STINNESBECK, W.,<br />

HOTTINGER, L., STUEBEN, D., BERNER, Z., RAMANIVOSOA, B. and<br />

RANDRIAMANANTENASOA, A. (2002) Age and Paleoenvironment<br />

of <strong>the</strong> Maastrichtian-Paleocene of <strong>the</strong> Mahajanga Basin,<br />

Madagascar: a multi-disciplinary approach. Mar. Micropaleo.,<br />

v.47, pp.17-70.<br />

ABRAMOVICH, S., KELLER, G., STUEBEN, D. and BERNER, Z. (2003)<br />

Characterization of late Campanian and Maastrichtian planktic<br />

foraminiferal depth habitats and vital activities based on stable<br />

iso<strong>to</strong>pes. Paleoclimat. Paleoecol. Paleogeog., v.202, pp.1-29.<br />

BAKSI, A.K. (2005) Comment on “ 40 Ar/ 39 Ar dating of <strong>the</strong><br />

Rajahmundry Traps, eastern India and <strong>the</strong>ir relations <strong>to</strong> <strong>the</strong><br />

<strong>Deccan</strong> Traps” by Knight et al. [Earth and Planet Sci. Lett.<br />

208 (2003) 85-99]. Earth Planet. Sci. Lett., v.239, pp.368-<br />

373.<br />

BAKSI, A.K., BYERLY, G.R. CHAN, L.-H. and FARRAR, E. (1994)<br />

Intracanyon flows in <strong>the</strong> <strong>Deccan</strong> province, India: Case his<strong>to</strong>ry<br />

of <strong>the</strong> Rajahmundry Traps. Geology, v.22, pp.605-608.<br />

BERGGREN, W.A., KENT, D.V., SWISHER, C.C. and AUBRY, M.-P.<br />

(1995) A revised Cenozoic geochronology and<br />

chronostratigraphy. In: W.A. Berggren, D.V. Kent, M.-P.<br />

Aubry, Nd J. Hardenbol (Eds.), Geochronology, Time Scales<br />

and Global Stratigraphic Correlation, SEPM, Spec. Publ.,<br />

no.54, pp.129-212.<br />

CANDE, S. and KENT, D.V. (1995) Revised calibration of <strong>the</strong><br />

geomagnetic polarity Timescale for <strong>the</strong> Late <strong>Cretaceous</strong> and<br />

Cenozoic. Jour. Geophys. Res., v.100, pp.6093-6095.<br />

CARON, M. (1985) <strong>Cretaceous</strong> planktic foraminifera. In: H.M. Bolli,<br />

J.B. Saunders and K. Perch-Nielsen (Eds.), Plank<strong>to</strong>n Stratigraphy:<br />

Cambridge. Cambridge University Press, pp.17-86.<br />

CHENET, A.-L., QUIDELLEUR, X., FLUTEAU, F. and COURTILLOT, V.<br />

(2007) 40 K/ 40 Ar dating of <strong>the</strong> main <strong>Deccan</strong> large igneous<br />

province: fur<strong>the</strong>r evidence of KTB age and short duration.<br />

Earth Planet. Sci. Lett., v.263, pp.1-15.<br />

CHENET, A.-L., FLUTEAU, F., COURTILLOT, V., GERARD, M. and<br />

SUBBARAO, K.V. (2008) Determination of rapid <strong>Deccan</strong><br />

eruptions across <strong>the</strong> <strong>Cretaceous</strong>-<strong>Tertiary</strong> boundary using<br />

paleomagnetic secular variation: Results from a 1200-m-thick<br />

section in <strong>the</strong> Mahabaleshwar. Jour. Geophys. Res., v.113,<br />

doi:10.1029/2006JB004635.<br />

CHENET, A.-L., COURTILLOT, V., FLUTEAU, F., GERARD, M.,<br />

QUIDELLEUR, X., KHADRI, S.F.R., SUBBARAO, K.V. and<br />

THORDARSON, T. (2009) Determination of rapid <strong>Deccan</strong><br />

eruptions across <strong>the</strong> <strong>Cretaceous</strong>-<strong>Tertiary</strong> boundary<br />

using paloemangnetic secular variation: 2. Constraints from<br />

analysis of eight new sections and syn<strong>the</strong>sis for a 3500 m-<br />

thick composite section. Jour. Geophys. Res., v.114, B06103,<br />

doi:10.1029/2008JB005644.<br />

COCCIONI, R. and LUCIANI, V. (2006) Guembelitria irregularis<br />

bloom at <strong>the</strong> K-T boundary: Morphological abnormalities<br />

induced by impact-related extreme environmental stress? In:<br />

C. Cockell, C. Koeberl and I. Gilmour, I. (Eds.), Biological<br />

Processes Associated with Impact Events. Impact Studies,<br />

Springer, pp.179-196.<br />

COURTILLOT, V., BESSE, J., VANDAMME, D., MONTIGNY, R., JAEGER,<br />

J.-J. and CAPPETTA, H. (1986) <strong>Deccan</strong> flood basalts at <strong>the</strong><br />

<strong>Cretaceous</strong>/<strong>Tertiary</strong> boundary? Earth Planet. Sci. Lett., v.80,<br />

pp.361-374.<br />

COURTILLOT, V., FERAUD, G., MALUSKI, H., VANDAMME, D., MOREAU,<br />

M.G. and BESSE, J. (1988) The <strong>Deccan</strong> flood basalts and <strong>the</strong><br />

<strong>Cretaceous</strong>-<strong>Tertiary</strong> boundary. Nature, v.333, pp.843-846.<br />

COURTILLOT, V., GALLET, Y., ROCCHIA, R., FERAUD, G., ROBIN, E.,<br />

HOFMANN, C., BHANDARI, N. and GHEVARIYA, Z.G. (2000)<br />

Cosmic markers, 40 Ar/ 39 Ar dating and Paleomagnetism of <strong>the</strong><br />

KT sections in <strong>the</strong> Anjar area of <strong>the</strong> <strong>Deccan</strong> large igneous<br />

province. Earth Planet. Sci. Lett., v.182, pp.137-156.<br />

COWIE, J.W., ZIEGLER, W. and REMANE, J. (1989) Stratigraphic<br />

Commission accelerates Progress, 1984 <strong>to</strong> 1989. Episodes,<br />

v.12, pp.79-83.<br />

DECONTO, R. M., BRADY, E.C., BERGENGREN, J. and HAY, W.W.<br />

(2000) Late <strong>Cretaceous</strong> climate, vegetation, and ocean<br />

interactions, in Warm Climate in Earth His<strong>to</strong>ry, In: B.T. Huber,<br />

K.G. MacLeod and S.L. Wing (Eds.), Cambridge Univ. Press,<br />

New York. pp.241-274.<br />

DESSERT, C., DUPRE, B., GAILLARDET, J., FRANCOIS, L.M. and<br />

ALLEGRE, C.J. (2003) Basalt wea<strong>the</strong>ring laws and <strong>the</strong> impact<br />

of basalt wea<strong>the</strong>ring on <strong>the</strong> global carbon cycle. Chemical<br />

Geol., v.202, pp.257-273.<br />

DONNADIEU, Y., PIERREHUMBERT, R., JACOB, R. and FLUTEAU, F.<br />

(2006) Modelling <strong>the</strong> primary control of paleogeography<br />

on <strong>Cretaceous</strong> climate. Earth Planet. Sci. Lett., v.248, pp.426-<br />

437.<br />

DUNCAN, R.A. and PYLE, D.G. (1988) Rapid Eruption of <strong>the</strong> <strong>Deccan</strong><br />

Flood Basalts at <strong>the</strong> <strong>Cretaceous</strong>/<strong>Tertiary</strong> boundary. Nature,<br />

v.333, pp.841-843.<br />

GARG, R., KHOWAJA-ATEEQUZZAMAN and PRASAD, V. (2006)<br />

Significant dinoflagellate cyst biohorizons in <strong>the</strong> Upper<br />

<strong>Cretaceous</strong>-Palaeocene succession in <strong>the</strong> Khasi Hills,<br />

Meghalaya. Jour. Geol. Soc. India, v.67, pp.737-747.<br />

GERTSCH, B., KELLER, G., ADATTE, T. GARG, R. PRASAD, V.,<br />

FLEITMANN, D. and BERNER, Z. (2011) Environmental effects<br />

of <strong>Deccan</strong> volvanism across <strong>the</strong> <strong>Cretaceous</strong>-<strong>Tertiary</strong> transition<br />

in Meghalaya, India. Earth Planet. Sci. Lett., v., pp.310,<br />

pp.272-285<br />

GOVINDAN, A. (1981) Foraminifera from <strong>the</strong> infra- and intertrappean<br />

subsurface sediments of Narsapur Well-1 and age of <strong>the</strong> <strong>Deccan</strong><br />

JOUR.GEOL.SOC.INDIA, VOL.78, NOV. 2011


DECCAN VOLCANISM LINKED TO THE CRETACEOUS-TERTIARY BOUNDARY MASS EXTINCTION 427<br />

Trap flows. Proc. 9 th Indian Colloquium of Micropalaeon<strong>to</strong>logy<br />

and Stratigraphy, pp.81-93.<br />

GRADSTEIN, F., OGG, J. and SMITH, A. (2004) A Geologic Time Scale.<br />

Cambridge, U.K. Cambridge University Press, 598p.<br />

JAIPRAKASH, B.C., SINGH, J. and RAJU, D.S.N. (1993) Foraminiferal<br />

events across <strong>the</strong> K/T boundary and age of <strong>Deccan</strong> volcanism<br />

in Palakollu area, Krishna-Godavari Basin, India. Jour. Geol.<br />

Soc. India, v.41, pp.105-117.<br />

JAY, A.E. and WIDDOWSON, M. (2008) Stratigraphy, structure and<br />

volcanology of <strong>the</strong> south-east <strong>Deccan</strong> continental flood basalt<br />

province: implications for eruptive extent and volumes. Jour.<br />

Geol. Soc. London, v.165, pp.177-188.<br />

JAY, A.E., MAC NIOCAILL, C., WIDDOSON, M., SELF, S. and TUNER,<br />

W. (2009) New Palaeomagnetic data from <strong>the</strong><br />

Mahabaleshwar Plateau, <strong>Deccan</strong> Flood Basalt Province, India:<br />

Implications from <strong>the</strong> volcanostratigraphic architecture of<br />

Continental Flood Basalt Provinces. Jour. Geol. Soc. London,<br />

v.166, pp.1-12.<br />

KELLER, G. (1988) Extinctions, survivorship and evolution across<br />

<strong>the</strong> <strong>Cretaceous</strong>/<strong>Tertiary</strong> boundary at El Kef, Tunisia. Mar.<br />

Micropal., v.13, pp.239-263.<br />

KELLER, G. (1989) Extended <strong>Cretaceous</strong>/<strong>Tertiary</strong> boundary<br />

extinctions and delayed population change in planktic<br />

foraminifera from Brazos River, Texas. Paleoceanography,<br />

v.4(3), pp.287-332.<br />

KELLER, G. (2001) The end-<strong>Cretaceous</strong> Mass Extinction: Year 2000<br />

Assessment. Jour. Planet. Space Sci., v.49, pp.817-830.<br />

KELLER, G. (2003) Biotic effects of impacts and volcanism. Earth<br />

Planet. Sci. Lett., v.215, pp.249-264.<br />

KELLER, G. (2008) Impact Stratigraphy: old principle - new reality.<br />

Geol. Soc. America, Special Paper no.437, pp.147-178.<br />

KELLER, G. and LINDINGER, M. (1989) Stable Iso<strong>to</strong>pe, TOC and<br />

CaCO 3<br />

record across <strong>the</strong> <strong>Cretaceous</strong>/<strong>Tertiary</strong> <strong>Boundary</strong> at El<br />

Kef, Tunisia, Paleogeog., Paleoclimat., Paleoecol., v.73(3/4),<br />

pp. 243-265.<br />

KELLER, G. and ABRAMOVICH, S. (2009) Lilliput effect in late<br />

Maastrichtian planktic foraminifera: Response <strong>to</strong><br />

environmental stress. Palaeogeo. Palaeoclimat., Palaeoec.,<br />

v.284, pp.47-62.<br />

KELLER, G., D’HONDT, S.L. and VALLIER, T.L. (1983) Multiple<br />

microtectite horizons in upper Eocene marine sediments: no<br />

evidence for mass extinctions Science, v.221, pp.150-152.<br />

KELLER, G., LI, L. and MACLEOD, N. (1995) The <strong>Cretaceous</strong>/<strong>Tertiary</strong><br />

boundary stra<strong>to</strong>type section at El Kef, Tunisia: How<br />

catastrophic was <strong>the</strong> mass extinction? Paleogeog., Paleoclimat.,<br />

Paleoecol., v.119, pp.221-254.<br />

KELLER, G., ADATTE, T., STINNESBECK, W., LUCIANI, V., KAROUI, N.<br />

and ZAGHBIB-TURKI, D. (2002) Paleoecology of <strong>the</strong> <strong>Cretaceous</strong>-<br />

<strong>Tertiary</strong> mass extinction in planktic foraminifera. Paleogeog.,<br />

Paleoclimat., Paleoecol., v.178, pp.257-298.<br />

KELLER, G., ADATTE, T., TANTAWY, A. A., BERNER, Z. and STUEBEN,<br />

D. (2007a) High Stress Late <strong>Cretaceous</strong> <strong>to</strong> early Danian<br />

paleoenvironment in <strong>the</strong> Neuquen Basin, Argentina. Cret. Res.,<br />

v.28, pp.939-960.<br />

KELLER, G., ADATTE, T., BERNER, Z., HARTING, M., BAUM, G., PRAUSS,<br />

M., TANTAWY, A.A. and STUEBEN, D. (2007b) Chicxulub impact<br />

predates KT boundary: new evidence from Brazos, Texas:<br />

Earth Planet. Sci. Lett., v.25(3-4), pp.339-356.<br />

KELLER, G., ADATTE, T., GARDIN, S., BARTOLINI, A. and BAJPAI, S.<br />

(2008) Main <strong>Deccan</strong> volcanism phase ends near <strong>the</strong> K-T<br />

boundary: Evidence from <strong>the</strong> Krishna-Godavari Basin, SE<br />

India. Earth Planet. Sci. Lett., v.268, pp.293-311.<br />

KELLER, G., KHOSLA, S.C., SHARMA, R., KHOSLA, A., BAJPAI, S. and<br />

ADATTE, T. (2009a) Early Danian planktic foraminifera from<br />

<strong>Cretaceous</strong>-<strong>Tertiary</strong> intertrappean beds at Jhilmili, Chhindwara<br />

District, Madhya Pradesh, India. Jour. Foram. Res., v.39,<br />

pp.40-55.<br />

KELLER, G., ADATTE, T., BAJPAI, S., MOHABEY, D.M., WIDDOWSON,<br />

M., KHOSLA, A., SHARMA, R., KHOSLA, S. C., GERTSCH, B.,<br />

FLEITMANN, D. and SAHNI, A. (2009b) K-T transition in <strong>Deccan</strong><br />

traps and intertrappean beds in central India mark major marine<br />

Seaway across India. Earth Planet. Sci. Lett., v.282, pp.10-<br />

23.<br />

KELLER, G., SAHNI, A. and BAJPAI, S. (2009c) <strong>Deccan</strong> volcanism,<br />

<strong>the</strong> KT mass extinction and dinosaurs. Jour. Biosciences, v.34,<br />

pp.709-728.<br />

KELLER, G., ADATTE, T., PARDO JUEZ, A. and LOPEZ-OLIVA, J. (2009d)<br />

New evidence concerning <strong>the</strong> age and biotic effects of <strong>the</strong><br />

Chicxulub impact in NE Mexico. Jour. Geol. Soc. London,<br />

v.166, pp.393-411.<br />

KELLER, G., ABRAMOVICH, S., BERNER, Z. and ADATTE, T. (2009e)<br />

Biotic effects of <strong>the</strong> Chicxulub impact, K-T catastrophe and<br />

sea-level change in Texas. Paleogeog. Paleoclimat. Paleoecol.,<br />

v.271, pp.52-68.<br />

KIDDER, K.L. and WORSLEY, T.R. (2010) Phanerozoic Large Igneous<br />

Provinces (LIPs), HEATT (Haline Euxinic Acidic Thermal<br />

Transgression) episodes, and mass extinctions: Palaeogeog.,<br />

Palaeoclimat., Palaeoecol., v.295, pp.162-191.<br />

KNIGHT, K.B. RENNE, P.R. HALKETT, A. and WHITE, N. (2003) 40 Ar/<br />

39 Ar dating of <strong>the</strong> Rajahmundry Traps, eastern India and <strong>the</strong>ir<br />

relationship <strong>to</strong> <strong>the</strong> <strong>Deccan</strong> Traps. Earth Planet. Sci. Lett.,<br />

v.208, pp. 85-99.<br />

KNIGHT, K.B., RENNE, P.R., BAKER, J., WAIGHT, T. and WHITE, N.<br />

(2005) Reply <strong>to</strong> 40 Ar/ 39 Ar dating of <strong>the</strong> Rajahmundry Traps,<br />

Eastern India and <strong>the</strong>ir relationship <strong>to</strong> <strong>the</strong> <strong>Deccan</strong> Traps:<br />

Discussion’ by A.K. Baksi. Earth Planet. Sci. Lett., v.239, pp.<br />

374-382.<br />

KUCERA, M. and MALMGREN, B.A. (l998) Terminal <strong>Cretaceous</strong><br />

warming event in <strong>the</strong> mid-latitude South Atlantic Ocean:<br />

evidence from poleward migration of Contusotruncana<br />

contusa (plank<strong>to</strong>nic foraminifera) morphotypes. Palaeogeog.,<br />

Palaeoclimat., Palaeoec, v.138, pp.1-15.<br />

LI, L. and KELLER, G. (1998a) Maastrichtian climate, productivity<br />

and faunal turnovers in planktic foraminifera of South Atlantic<br />

DSDP Sites 525A and 21. Mar. Micropal., v.33(1-2), pp. 55-<br />

86.<br />

LI, L. and KELLER, G. (l998b) Diversification and extinction in<br />

Campanian-Maastrichtian planktic foraminifera of<br />

northwestern Tunisia: Eclogae Geol. Helvetiae, v.91, pp.75-<br />

102.<br />

JOUR.GEOL.SOC.INDIA, VOL.78, NOV. 2011


428 G. KELLER AND OTHERS<br />

LI, L. and KELLER, G. (1998c) Abrupt deep-sea warming at <strong>the</strong> end<br />

of <strong>the</strong> <strong>Cretaceous</strong>. Geology, v.26(11), pp.995-998.<br />

LUCIANI, V. (2002) High resolution plank<strong>to</strong>nic foraminiferal<br />

analysis from <strong>the</strong> <strong>Cretaceous</strong>/<strong>Tertiary</strong> boundary at Ain Settara<br />

(Tunisia): Evidence of an extended mass extinction.<br />

Palaeogeog., Palaeoclimat., Palaeoec., v.178 (3), pp.299-319.<br />

MALARKODI, N., KELLER, G., FAYAZUDEEN, P.J. and MALLIKARJUNA,<br />

U.B. (2010) Foraminifera from <strong>the</strong> early Danian Intertrappean<br />

beds in Rajahmundry Quarries, Andhra Pradesh, SE India.<br />

Jour. Geol. Soc. India, v.75, pp.851-863.<br />

MCLEAN, D. (1978)A terminal Mesozoic “Greenhouse”: Lessons<br />

from <strong>the</strong> past. Science, v.201 (4354), pp.401-406.<br />

MCLEAN, D. (1985) <strong>Deccan</strong> Traps Mantle Degassing in <strong>the</strong> Terminal<br />

<strong>Cretaceous</strong> Marine Extinctions. <strong>Cretaceous</strong> Res., v.6, pp.235-<br />

259.<br />

MEHROTRA, N.C. and SARGEANT, W.A.S. (1987) Late <strong>Cretaceous</strong><br />

<strong>to</strong> Early <strong>Tertiary</strong> dinoflagellate cysts from Narsapur Well -1,<br />

Godavari-Krishna Basin, south-India, Geobios, v.20, pp.149-<br />

191.<br />

MISRA, K.S. (2005) Distribution pattern, age and duration and mode<br />

of eruption of <strong>Deccan</strong> and associated volcanics. Gondwana<br />

Geol. Mag., v.8, pp.53-60.<br />

Nordt, L., Atchley, S. and Dworkin, S. (2003) Terrestrial Evidence<br />

for Two Greenhouse Events in <strong>the</strong> Latest <strong>Cretaceous</strong>. GSA<br />

Today, v.13(12), pp.4-9.<br />

OLSSON, R.K., WRIGHT, J.D. and MILLER, K.D. (2001)<br />

Palobiogeography of Pseudotextularia elegans during <strong>the</strong><br />

latest Maastrichtian global warming event. Jour. Foram. Res.,<br />

v.31, pp.275-282.<br />

PANDE, K., PATTANAYAK, S.K., SUBBARAO, K.V.,<br />

NAVANEETHAKRISHNAN, P. and VENKATESAN, T.R. (2004) 40 Ar/<br />

39 Ar age of a lava flow from <strong>the</strong> Bhimashankar Formation,<br />

Giravali Ghat, <strong>Deccan</strong> Traps. Proc. Indian Acad. Sci., v.113,<br />

pp.755–758.<br />

PARDO, A. and KELLER, G. (2008) Biotic effects of environmental<br />

catastrophes at <strong>the</strong> end of <strong>the</strong> <strong>Cretaceous</strong>: Guembelitria and<br />

Heterohelix blooms. <strong>Cretaceous</strong> Res., v.29 (5/6), pp.1058-<br />

1073.<br />

PRASAD, B. and PUNDEER, B.S. (2002) Palynological events and<br />

zones in <strong>Cretaceous</strong>- <strong>Tertiary</strong> <strong>Boundary</strong> sediments of Krishna-<br />

Godavari and Cauvery basins, India. Palaeon<strong>to</strong>graphica Abt.<br />

B, v.262, pp.39-70.<br />

RAJU, D.S.N. (2008) Stratigraphy of India. ONGC Bull., Spec.<br />

Issue, no.43(1), 44p.<br />

RAJU, D.S.N., JAIPRAKASH, B.C., KUMAR, A., SAXENA, R.K., DAVE,<br />

A., CHATTERJEE, T.K. and MISHRA, C.M. (1994) The magnitude<br />

of hiatus and sea-level changes across K/T boundary in<br />

Cauvery and Krishna-Godavari basins. India. Jour. Geol. Soc.<br />

India, v.44, pp.301-315.<br />

RAJU, D.S.N., JAIPRAKASH, B.C. KUMAR, A. SAXENA, R.K. DAVE, A.<br />

CHATTERJEE, T.K. and MISHRA, C.M. (1995) Age of <strong>Deccan</strong><br />

volcanism across KTB in Krishna-Godavari Basin: new<br />

evidences. Jour. Geol. Soc. India, v.45, pp.229-233.<br />

RAJU, D.S.N. JAIPRAKASH, B.C. and KUMAR, A. (1996)<br />

Paleoenvironmental set-up and age of basin floor just prior <strong>to</strong><br />

<strong>the</strong> spread of <strong>Deccan</strong> volcanism in <strong>the</strong> Krishna-Godavari Basin,<br />

India. Mem Geol. Soc. India, v.37, pp.285-295.<br />

REMANE, J., KELLER, G., HARDENBOL, J. and BEN HAJ ALI, M. (l999)<br />

Report on <strong>the</strong> international workshop on <strong>Cretaceous</strong>-Paleogene<br />

transition: Episodes, v.22(1), pp.47-48.<br />

ROCCHIA, R., ROBIN, E., FROGET, L. and GAYRAUD, J. (1996)<br />

Stratigraphic distribution of extraterrestrial markers at <strong>the</strong><br />

<strong>Cretaceous</strong>-<strong>Tertiary</strong> boundary in <strong>the</strong> Gulf of Mexico area:<br />

implications for <strong>the</strong> temporal complexity of <strong>the</strong> event. In:<br />

Ryder, G., Fas<strong>to</strong>vsky, D. and Gartner, S. (Eds.), The<br />

<strong>Cretaceous</strong>-<strong>Tertiary</strong> event and o<strong>the</strong>r Catastrophes in Earth<br />

his<strong>to</strong>ry. Geol. Soc. America, Spec. Paper no.307, pp.279-286.<br />

SAXENA, R.K. and MISRA, C.M. (1994) Time and duration of <strong>Deccan</strong><br />

volcanism in <strong>the</strong> Razole area, Krishna-Godavari Basin, India.<br />

Curr. Sci., v.66(1), pp.73-76.<br />

SAXENA, R.K. and MISRA, C.M. (1995) Campanian-Maastrichtian<br />

nannoplank<strong>to</strong>n biostratigraphy of <strong>the</strong> Narsapur clays<strong>to</strong>ne<br />

Formation, Krishna-Godari Basin, India. Jour. Geol. Soc. India,<br />

v.45, pp.323-329.<br />

SELF, S., JAY, A.E., WIDDOWSON, M. and KESZTHELYI, L.P. (2008a)<br />

Correlation of <strong>the</strong> <strong>Deccan</strong> and Rajahmundry Trap lavas: Are<br />

<strong>the</strong>se <strong>the</strong> longest and largest lava flows on Earth? Jour. Volc.<br />

and Geoth. Res., v.172, pp.3-19.<br />

SELF, S., BLAKE, S., SHARMA, K., WIDDOWSON, M. and SEPHTON, S.<br />

(2008b) Sulfur and Chlorine in Late <strong>Cretaceous</strong> <strong>Deccan</strong><br />

Magmas and Eruptive gas release. Science, v.319, pp.1654-<br />

1657.<br />

SHETH, H.C., PANDE, K. and BHUTANI, R. (2001) 40 Ar/ 39 Ar age of a<br />

national geological monument: The gilbert Hill basalt, <strong>Deccan</strong><br />

Traps, Bombay. Curr. Sci., v.80, pp.1437-1440.<br />

TANTAWY A.A. (2003) Calcareous nannofossil biostratigraphy and<br />

paleoecology of <strong>the</strong> <strong>Cretaceous</strong>-<strong>Tertiary</strong> transition in <strong>the</strong><br />

western desert of Egypt. Mar. Micropal., v.47, pp.323-356.<br />

VENKATESAN, T.R., PANDE, K. and GOPALAN, K. (1993) Did <strong>Deccan</strong><br />

volcanism pre-date <strong>the</strong> <strong>Cretaceous</strong>-<strong>Tertiary</strong> transition? Earth<br />

Planet. Sci. Lett., v.119(1-2), pp.181-189.<br />

VENKATARENGAN, R., RAO, G.N., PRABHAKAR, K.N., SINGH, D.N.,<br />

AWASTHI, A.K., REDDY, P.K., MISHRA, P.K. and ROY, S.K. (1993)<br />

Lithostratigraphy of <strong>the</strong> Krishna-Godavari Basin, Unpublished<br />

ONGC Internal Document III, Allied Printers, Dehradun, pp.1-<br />

29.<br />

VON SALIS, K. and SAXENA, R.K. (1998) Calcareous nannofossils<br />

across <strong>the</strong> K/T boundary and <strong>the</strong> age of <strong>the</strong> <strong>Deccan</strong> Trap<br />

volcanism in sou<strong>the</strong>rn India. Jour. Geol. Soc. India, v.51,<br />

pp.183-192.<br />

WILF, P., JOHNSON, K.R. and HUBER, B.T. (2003) Correlated<br />

terrestrial and marine evidence for global climate changes<br />

before mass extinction at <strong>the</strong> <strong>Cretaceous</strong>-Paleogene boundary.<br />

Proc. Nat. Acad. Sci., USA, v.100(2), pp.599-604.<br />

(Received: 20 May 2011;Revised form accepted: 9 June 2011)<br />

JOUR.GEOL.SOC.INDIA, VOL.78, NOV. 2011

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!