Вы находитесь на странице: 1из 29

Downloaded from specialpapers.gsapubs.

org on March 18, 2011

Geological Society of America Special Papers


Cretaceous paleoceanography: Evidence from planktonic foraminiferal evolution
Isabella Premoli Silva and William V. Sliter Geological Society of America Special Papers 1999;332;301-328 doi: 10.1130/0-8137-2332-9.301

Email alerting services Subscribe Permission request

click www.gsapubs.org/cgi/alerts to receive free e-mail alerts when new articles cite this article click www.gsapubs.org/subscriptions/ to subscribe to Geological Society of America Special Papers click http://www.geosociety.org/pubs/copyrt.htm#gsa to contact GSA

Copyright not claimed on content prepared wholly by U.S. government employees within scope of their employment. Individual scientists are hereby granted permission, without fees or further requests to GSA, to use a single figure, a single table, and/or a brief paragraph of text in subsequent works and to make unlimited copies of items in GSA's journals for noncommercial use in classrooms to further education and science. This file may not be posted to any Web site, but authors may post the abstracts only of their articles on their own or their organization's Web site providing the posting includes a reference to the article's full citation. GSA provides this and other forums for the presentation of diverse opinions and positions by scientists worldwide, regardless of their race, citizenship, gender, religion, or political viewpoint. Opinions presented in this publication do not reflect official positions of the Society.

Notes

Geological Society of America

Downloaded from specialpapers.gsapubs.org on March 18, 2011

Geological Society of America Special Paper 332 1999

Cretaceous paleoceanography: Evidence from planktonic foraminiferal evolution


Isabella Premoli Silva Dipartimento di Scienze della Terra, Universita di Milan, 20133 Milano, Italy William V. Sliter (deceased October 31, 1997) U.S. Geological Survey, 345 Middlefield Road, MS 975, Menlo Park, California 94025-3591

ABSTRACT The evolution of planktonic foraminifers in the Cretaceous shows pulses of diversification and stasis interrupted by brief extinction events and faunal turnover. The overall record shows a threefold pattern; from the early Valanginian to the latest Aptian/Albian boundary, then to the latest Albian, and finally, to the end of the Cretaceous. The pattern of evolution in the first and second intervals is similar and is characterized by increasing diversity, size, and morphologic complexity. The third interval differs in pattern and shows short periods of rapid diversification and turnover separated by longer periods of stasis. We equate these evolutionary changes to parallel changes in the physical and chemical structure of the Cretaceous ocean. The first and second intervals represent the progression from a mixed, eutrophic, upper-water column to the development of a thermocline, stratification, and nichepartitioning in an oligotrophic column. The third interval is more complex and is characterized by periods of mixing and stability above the thermocline probably controlled by climatic conditions and nutrient runoff, and by the development of welldefined latitudinal bioprovinces in the Campanian and Maastrichtian. This overall record is interrupted by five events of paleoceanographic significance. Three of the events, the Selli Event in the early Aptian, the Aptian/Albian boundary event, and the Bonarelli Event in the latest Cenomanian, are defined by the deposition of organic carbon-rich sediment. The fourth, or Santonian event, precipitated the largest foraminiferal turnover during the Cretaceous, affecting all planktonic foraminiferal trophic groups. This event also ushered in a Cretaceous ocean with modern attributes. The fifth event was the catastrophic extinction at the end of the Cretaceous that terminated the Mesozoic Era. Each of these events shows evidence of upper-water column disruption likely related to increased upwelling.

INTRODUCTION Modern planktonic foraminiferal distributions define five latitudinally constrained bioprovinces from the equator to the poles. Altogether, the bioprovinces are characterized by a marked decrease in species richness (or diversity) from the equator toward the high latitudes where the latter assemblages are almost

monospecific and are dominated by the most opportunistic species with simple morphologies (B, 1982). Similar patterns of distribution can be reconstructed as far back in time as the Paleocene, although several of the Paleogene Epoch morphologic groups are extinct and apparently did not behave like their modern counterparts (Boersma and Premoli Silva, 1983, 1991). How do the modern, or at least the Paleogene, patterns apply to Creta-

Premoli Silva, I., and Sliter, W. V., 1999, Cretaceous paleoceanography: Evidence from planktonic foraminiferal evolution, in Barrera, E., and Johnson, C. C., eds., Evolution of the Cretaceous Ocean-Climate System: Boulder, Colorado, Geological Society of America Special Paper 332.

301

Downloaded from specialpapers.gsapubs.org on March 18, 2011

302

I. Premoli Silva and W. V. Sliter the prolonged stasis (about 10 m.y.) during most of the Valanginian and Hauterivian after the first diversification, whereas the following periods of diversification occurred over a comparable length of time in both intervals. The third interval, from near the close of the Albian until the end of the Cretaceous, differs in pattern from the earlier intervals (Fig. 1). This interval is characterized by short periods of rapid diversification and turnover separated by longer periods of stasis except around the Corg-rich Bonarelli Event near the Cenomanian/Turonian boundary (Arthur et al., 1990) when the cycles of alternation were strongly accelerated (two periods of diversification separated by a stasis within <6 m.y.). Paleoceanographically, we believe the first and second intervals show a similar evolution of the upper water column that progressed from a mixing stage to a more stable stage (Fig. 1). In both cases, this evolution led to the development of a thermocline that was weakly developed in the first interval but was well defined by the second. Moreover, after the episode of strong mixing around the Aptian/Albian boundary, which separates the first and second intervals, the early to middle Albian ocean experienced a return to the Hauterivian state characterized by gradual diversification and closely spaced fluctuations in planktonic foraminiferal abundance, suggesting a linkage to Milankovitch orbital cycles. This state was followed in the middle Albian by paleoceanographic conditions that replicated the late Aptian state characterized by rapid diversification and test modifications in the planktonic foraminifers. The onset of a strong thermocline in the late Albian heralded for the first time in geological history the development of oceanic conditions similar to those that characterize the modern ocean. This first stratification of the Cretaceous water masses can be explained by an increase in the Earths latitudinal thermal gradient, which resulted in the differentiation of discrete latitudinal bioprovinces. Within this paleoceanographic scenario, the third interval began with the latitudinal expansion of warm, dry, climatic conditions in the presence of a well-established, deep thermocline. This early Cenomanian state was disrupted by the Bonarelli Event, which is characterized by strong mixing during warm climatic conditions. Notably, evidence of increased mixing accompanied by a weakening of the thermocline in the upper-water column in the middle to late Cenomanian preceded the Bonarelli Event, that we interpret as a short disruption of the thermocline (Arthur and Premoli Silva, 1982; Premoli Silva and Sliter, 1994). Paleoceanographic conditions replicating the Cenomanian state rapidly resumed during the Turonian, soon after the Bonarelli Event. Fluctuating climatic conditions probably controlled nutrient distribution in the surface waters and governed the oceanic regime in the later part of the Cretaceous. This along with the increase of latitudinal and vertical temperature gradients during the Campanian led to the differentiation of discrete bioprovinces similar to those in the modern ocean. Again the trend was interrupted, this time by the terminal Cretaceous Event that, as with

ceous planktonic foraminifers that are extinct except for the opportunistic Guembelitria? Can the evolutionary changes displayed by Cretaceous planktonic foraminifers be used to interpret the paleoceanographic evolution of the Cretaceous ocean? Here, we summarize the state of the art and interpret the sequence of major ecological changes in the pelagic realm during the Cretaceous based primarily on the evolutionary development (morphologic complexity, diversity, life strategy) and spatial distribution of planktonic foraminifers, from their first diversification in the early Valanginian to the crisis at the Cretaceous/Paleogene Series boundary. Our intent is to extend our paleoceanographic interpretation to the fullest to foster future discussion and comparative analysis. While nowhere rigorous, we also incorporate selected significant changes shown by other fossil groups such as calcareous nannofossils, radiolarians, dinoflagellates, benthic foraminifers, and megafossils to augment our conclusions. Earlier preliminary paleoceanographic reconstructions were based on Albian to Maastrichtian planktonic foraminiferal faunas from the remarkably continuous pelagic record from central Italy (Wonders, 1980; Premoli Silva and Sliter, 1994). Here, we extend the record through the Early Cretaceous by incorporating the results of paleontologic investigations from the Rio Argos section in southern Spain (Coccioni and Premoli Silva, 1994) and the Gorgo a Cerbara section in Italy (Coccioni et al., 1992; Cecca et al., 1994). The Bottaccione and Rio Argos sections together span the entire evolutionary succession of Cretaceous planktonic foraminifers. Both sections are well calibrated to integrated bioand magnetostratigraphic schemes and to most stages within the latest chronologic time scale (Erba et al., 1995b, mainly after Gradstein et al., 1994; Channell et al., 1995). CRETACEOUS PALEOCEANOGRAPHIC CHANGES The record of planktonic foraminiferal evolution through the Cretaceous is characterized by periods of diversification that alternate with periods of stasis during which the assemblages apparently underwent little or no change (Fig. 1). These alternating periods of slow to rapid diversification and stasis display different duration through time. We equate these evolutionary patterns to parallel changes in the physical and chemical structure of the worlds oceanic water masses. The overall Cretaceous record shows a general threefold pattern. The first interval extends from the first diversification in the early Valanginian until the latest Aptian. The pattern shows a continuously increasing diversification that is interrupted by a single episode of moderate turnover near the Selli Event, an episode of organic carbon-rich (Corg-rich) sediment deposition in the late early Aptian (Coccioni et al., 1992; Erba, 1994). The second interval extends from the Aptian/Albian boundary, another Corgrich episode (Brhret et al., 1986; Tornaghi et al., 1989; Arthur et al., 1990), and extends until the latest Albian. Again the pattern is characterized by continuously increasing diversification similar to that of the first interval but in half of the time (12 m.y. versus 22 m.y.). The difference in duration is mainly related to

Downloaded from specialpapers.gsapubs.org on March 18, 2011

Figure 1. Major planktonic foraminiferal evolutionary patterns and paleoceanographic changes through the Cretaceous plotted against planktonic foraminiferal zonal scheme and major stratigraphic events, magnetostratigraphy, and absolute age (from Erba et al., 1995b; and Channell et al., 1995). Arrows for evolutionary changes not in scale and exaggerated for changes in diversification. P, precursor events; T, turnover; E, extinctions.

Downloaded from specialpapers.gsapubs.org on March 18, 2011

304

I. Premoli Silva and W. V. Sliter a decrease in the early Campanian, species diversity then increased continuously from the late early Campanian until the Gansserina gansseri Zone near the base of the Maastrichtian, when there were more than 60 species, representing the highest diversification noted in the Cretaceous. After this peak, species diversity decreased progressively to 30 species just prior to the end of the Cretaceous. Within this overall trend, slow diversification and/or nondiversification characterize the upper part of the Dicarinella concavata Zone in the Coniacian and the middle part of the Globotruncanita elevata Zone in the early Campanian. The species diversity curve in general reflects the number of originations (Fig. 3). Major peaks in origination are recorded in the early Barremian Hedbergella similisHedbergella kuznetsovae Zone, at the base of the latest Albian Rotalipora appenninica Zone, in the upper part of the late Cenomanian Dicarinella algeriana Subzone of the R. cushmani Zone, in the lower part of the late Turonian D. primitivaMarginotruncana sigali Zone, throughout much of the lower and middle parts of the Santonian Dicarinella asymetrica Zone, at the base of the middle Campanian Globotruncana ventricosa Zone, and from the Globotruncana aegyptiaca to the base of Gansserina gansseri Zones in the latest Campanian. In contrast, extinctions contribute more significantly to the evolutionary overturn in the late Aptian, in the upper part of the latest Albian Rotalipora appenninica Zone, in the lower part of the D. asymetrica Zone, in the lower part of the early Campanian Globotruncanita elevata Zone, and in the middle to latest Maastrichtian. Overall throughout the Cretaceous, however, originations predominate over extinctions. The curve for the degree of turnover shown in Figure 3, which was derived by adding the number of originations and extinctions, is more significant than the separate curves of the latter two criteria. In particular, two peaks stand out, one in the Santonian Dicarinella asymetrica Zone and a second, larger one just prior to the end of the Cretaceous. Smaller but discrete peaks are also recorded in the early Barremian, early Aptian, latest Albian, late Cenomanian, late middle Turonian, and latest Campanian. All significant peaks in turnover are expressed schematically but not to scale in Figure 1 by the larger inverted triangles in the column for planktonic foraminiferal changes. MODERN PLANKTONIC FORAMINIFERAL LIFE STRATEGIES Here we examine the life strategies of modern planktonic foraminifers as a model for our interpretation of the evolution of Cretaceous species. Most modern planktonic foraminifers live discretely stratified within the upper part of the water column above the thermocline, in the so-called mixed layer (Hemleben et al., 1989). Because the thermocline is typically deeper in low-latitude tropical regions and shallows toward high latitudes where it reaches the surface at the circumpolar front, the volume of the mixed layer decreases with increased latitude, resulting in the progressive elimination of niches inhabited by planktonic

the Bonarelli Event in the latest Cenomanian, was preceded by increased oceanic mixing and closely spaced temperature fluctuations possibly controlled by Milankovitch orbital cycles (see Herbert et al., 1995; Streel et al., 1995). Below, we examine the basis of our paleoceanographic conclusions derived from the evolution and distribution patterns of Cretaceous planktonic foraminifers and other selected fossil groups as well as generalized information from sedimentary and isotopic records. PATTERNS OF EVOLUTIONARY CHANGES IN CRETACEOUS PLANKTONIC FORAMINIFERS The evolutionary patterns exhibited by Cretaceous planktonic foraminifers are based on the records of species and generic richness determined from published biostratigraphic ranges (e.g., Sigal, 1977; Wonders, 1979, 1980; Robazynski et al., 1984; Caron, 1985; Huber, 1990, 1991; Sliter, 1992; Coccioni and Premoli Silva, 1994; Premoli Silva and Sliter, 1994; Robazynski and Caron, 1995). The derived values are plotted graphically in Figure 2 against the planktonic foraminiferal zonal scheme and time scale of Erba et al. (1995b, mainly after Gradstein et al., 1994), and updated in the Early Cretaceous according to new ages provided by Channell et al. (1995). The genera and species used in our computations are listed in Appendix A. Our analyses do not include the near-shore favusellids or globuligerinids in the Early Cretaceous, species recently described as endemic to the Antarctic region, or those recorded only at one location (e.g., Haig, 1992; Huber, 1988; Nederbragt, 1991). Consequently, species richness is generally underestimated. Figure 2 clearly shows the diversity of planktonic foraminifers increasing overall at both the generic and specific levels from the first diversification in the early Valanginian to the end of the Cretaceous. The relative number of species generally parallels the number of genera (see also Premoli Silva and Sliter, 1994), however, the plot of species diversity (Fig. 3) shows that species richness increased discontinuously throughout the Cretaceous. For example, the Early Cretaceous Period is characterized by a slow gradual diversification from the early Valanginian to the base of the Globigerinelloides ferreolensis Zone in the early late Aptian. This diversification is followed by a progressive but relatively rapid decrease in species richness during the remainder of the Aptian until very low numbers were reached in the interval around the Aptian/Albian boundary. Species diversification then slowly rebounded during most of the Albian, with an acceleration in the late Albian Rotalipora appenninica Zone. After a slight decrease in the latest Albian and early Cenomanian, a gradual increase in diversity characterized the remainder of the Cenomanian and Turonian. The exception in this pattern occurs at the Bonarelli Event; a short episode characterized by low diversity within the planktonic foraminifers balanced against a high abundance of heterohelicids and Hedbergella planispira. Diversity decreased slightly in the Coniacian and then peaked in the Santonian Dicarinella asymetrica Zone. Following

Downloaded from specialpapers.gsapubs.org on March 18, 2011

Cretaceous paleoceanography: Evidence from planktonic foraminiferal evolution

305

Figure 2. Stratigraphic distribution of planktonic foraminiferal genera and number of species per genus plotted against planktonic foraminiferal zonal scheme, magnetostratigraphy, and absolute age (as in Fig. 1). Compilation based mainly on Pessagno (1967), Longoria (1974), Robaszynski et al. (1979, 1984, 1990, 1993), Caron (1985), Coccioni and Premoli Silva (1994), Premoli Silva and Sliter (1994), and Robaszynski and Caron (1995). Columns on the right show the timing of the five events that interrupt the Cretaceous planktonic foraminiferal record and the broad threefold evolutionary pattern.

foraminifers. As a consequence, planktonic foraminifers inhabiting the mixed layer characteristically decrease in diversity from the tropics toward the high latitudes and in general are absent in polar waters (B, 1977). Accompanying the latitudinal decrease in species richness from tropics to high latitude is the progressive loss of the less-tolerant planktonic foraminiferal species, which are characterized by complex morphologies and/or high-energy requirements. As

a result, high-latitude assemblages become dominated by the most tolerant cosmopolitan forms that are characterized by simple morphologies (Margalef, 1965; B, 1980, 1982; Hemleben et al., 1989). Similar decreases in morphologic complexity are observed in planktonic foraminiferal assemblages associated with the inception and intensification of seasonal upwelling regimes (Kroon and Ganssen, 1988). Further, laboratory experiments and the results of plankton

Downloaded from specialpapers.gsapubs.org on March 18, 2011

Figure 3. Number of species, originations, extinctions, and degree of turnover in Cretaceous planktonic foraminifers (foraminiferal sources as in Fig. 2, chronology as in Fig. 1). Degree of turnover is the sum of the number of originations and extinctions.

Downloaded from specialpapers.gsapubs.org on March 18, 2011

Cretaceous paleoceanography: Evidence from planktonic foraminiferal evolution tows from vertical transects in key areas of the oceans (see Hemleben et al., 1989, and references therein) show that the vertical and horizontal distribution and abundance of planktonic foraminifers are controlled by numerous biotic and abiotic factors, which can enhance or limit population growth. Among these are reproductive capacity, presence or absence of symbionts, diversity of nutritional sources, and varied physical and chemical parameters. Although we are far from fully understanding these interacting relationships, we can state the following: 1. The different depths occupied by planktonic foraminifers in the upper water column likely reflect the avoidance of competition and the efficient exploitation of unique conditions prevailing in each water horizon. 2. Maximum vertical separation of species occurs in warm waters of tropical and subtropical regions that are characterized by greater diversity of physical and biotic variables with depth. Vertical separation is much less evident, if present, in cooler waters, which become more uniform in structure with increased latitude. 3. The wide range of niches inhabited by planktonic foraminifers in tropical waters is due to large variations in salinity and temperature in the upper few hundred meters; this correlates to variations in abundance and diversity of potential prey, diversified habitats, and multiple adaptive strategies. The result is a rich species diversity. 4. The various preferred depth habitats are characterized by distinct changes in morphology, surface ornamentation, possession/absence of symbionts, and reproductive cycle. In general, globigerinids, especially spinose species, mainly abound in surface waters, whereas a number of nonspinose species, including globorotaliid morphologies, are found at depth, although some may be abundant in temperate surface waters during the winter months. 5. Shell calcification occurs mainly in the upper hundred meters, although there is evidence for continuing calcification, thus growth, while sinking to greater depths (deeper than 100 m) and even below the thermocline (i.e., Globorotalia menardii, Truncorotalia truncatulinoides, Globorotalia hirsuta, Neogloboquadrina dutertrei, among others). 6. Reproduction occurs with a lunar or semilunar periodicity in several species. 7. Gamete release occurs at the thermocline at the base of the mixed layer. The thermocline is a stable layer that coincides with the deep chlorophyll maximum (DCM) layer and is rich in phytoplankton. Most juveniles of spinose and nonspinose species that require phytoplankton prey are found in the thermocline region even if, as adults, some of these switch to a more carnivorous diet (Fairbanks and Wiebe, 1980; Hemleben et al., 1989). 8. Surface-dwelling spinose species show a distinct preference for zooplankton prey. In contrast, nonspinose species, which include most of the deep dwellers, show a tendency to consume phytoplankton prey. Consequently, phytoplankton distribution and composition influence the spatial and temporal distribution of each species.

307

9. Some species with complex morphologies (i.e., T. truncatulinoides and G. hirsuta) apparently reproduce in annual cycles at locations where spring blooms provide ample phytoplankton prey and water temperatures are suitable for survival. During the rest of the year they live in cooler, deeper water in tropical areas, probably feeding on decaying organic matter and marine snow, but move progressively closer to the surface toward higher latitudes. These patterns emphasize the importance of temperature as a controlling factor for reproduction. 10. Spinose species are more abundant in oligotrophic waters, whereas nonspinose species are more abundant in eutrophic waters that are richer in phytoplankton (e.g., upwelling areas, diatom-rich Antarctic waters). 11. The latitudinal and vertical distribution of planktonic foraminiferal species is reflected in the oxygen and carbon isotopic ratios of their tests. Relationship to nutrients As reported by Hallock (1987), habitat diversity is inversely related to nutrients because (a) the length and complexity of food chains tend to be an inverse function of food supply, (b) eutrophic (nutrient-rich) environments are inherently unstable, and (c) the greater depths of low-nutrient (oligotrophic) euphotic zones, characterized by a range of light intensities, provide greater potential for specialization than do the shallower and more variable euphotic zones found in areas influenced by upwelling, runoff, or seasonal blooms. Some taxonomic groups, however, cannot take advantage of the potentially numerous niches of the low-nutrient environments because they metabolically require more energy than is available under such conditions. For instance, phytoplankton respond to nutrient supply as follows (Rhyther, 1969): diatoms bloom and flourish where upwelling and other sources of turbulence supply abundant nutrients, dinoflagellates succeed the diatoms as nutrient supplies decline and/or physical stability is established, and when nutrient supplies become scarce, nano- and picoplankton, including coccolithophorids, dominate the phytoplankton. This sequence of plankton communities is similar along a nutrient gradient in tropical-subtropical regions as well as through latitudes (Hallock, 1987). In oligotrophic subtropical gyres, complex food webs are nano- and picoplankton based (Rhyther, 1969). Dinoflagellates dominate areas or seasons having intermediate nutrient supplies, as in temperate waters, whereas low- and middle-latitude upwelling zones and high-latitude regions are dominated by diatoms (Hallock et al., 1991). Similarly, Globigerina bulloides, although the index species of the subpolar bioprovince (B, 1977; Kennett, 1982), is typically abundant in meridional upwelling zones in the tropics, even where water temperatures are depressed only a few degrees (De Mir, 1971). The similarity between low-latitude upwelling assemblages and those of high-latitude regions is easily explained in terms of energy requirements (Hallock et al., 1991). In organisms whose body temperatures conform closely with ambient temperatures,

Downloaded from specialpapers.gsapubs.org on March 18, 2011

308

I. Premoli Silva and W. V. Sliter nated by mesotrophic conditions. In these cases, taxa representing different water masses or ecologic conditions, ranging from relatively eutrophic during seasonal blooms to oligotrophic when upwelling or runoff decline seasonally or yearly, accumulate into the sediments. Diversity in these areas can also increase because of the presence of species unique to the ecotone. In summary, planktonic foraminifers inhabiting the mixed layer today characteristically decrease in number from the tropics toward the high latitudes and in general are absent in polar waters (B, 1980, 1982). The decrease in species richness from tropics to high latitude is reflected in the planktonic foraminiferal assemblages that progressively lose the less tolerant species (K-selected strategists), characterized by complex morphologies and/or highenergy requirements, and become dominated by the most tolerant, cosmopolitan, and opportunistic forms, characterized by small-sized simple morphologies (Margalef, 1965; B, 1980, 1982; Hemleben et al., 1989). Similar patterns characterize intensifying upwelling regimes (Kroon and Ganssen, 1988). Diversity decreases with increasingly less stable conditions in the upper water column. As a consequence, the decrease or disappearance of stratification in these areas results in a decrease in the number of ecologic niches. When conditions are extreme, the assemblages are dominated by a single species; Globigerina bulloides in the meridional upwelling areas (Kroon and Ganssen, 1988) and Neogloboquadrina pachyderma in the Polar bioprovince (B, 1977). LATITUDINAL AND VERTICAL DISTRIBUTION OF CRETACEOUS PLANKTONIC FORAMINIFERS We believe that Cretaceous planktonic foraminifers, although all extinct except for Guembelitria, behaved for the most part like their modern counterparts (Davids, 1966; Sliter, 1972; Wonders, 1980; Caron and Homewood, 1983; Caron, 1983; Leckie, 1989; Coccioni et al., 1992; Keller et al., 1993). This interpretation is supported by the few studies dealing with biogeography and isotopic analyses of Cretaceous planktonic foraminifers (e.g., Laughton et al., 1972; Hart and Tarling, 1974; Douglas and Savin, 1975; Hart and Bailey, 1979; Boersma and Shackleton, 1981; Huber, 1988, 1991; Haig, 1992; Huber et al., 1995) and is consistent with Paleogene reconstructions (Boersma and Premoli Silva, 1988, 1991; Premoli Silva and Boersma, 1989). At the same time, we acknowledge that differences existed between Cretaceous planktonic foraminifers and the modern stock, such as the fact that even the most simple Cretaceous morphotypes with globular chambers like the hedbergellids are finely perforated, do not bear spines, and probably lacked symbionts. Despite these differences, we believe that Cretaceous planktonic species with complex morphotypes preferentially inhabited tropical and subtropical oligotrophic waters, while high-latitude assemblages were dominated by small-sized opportunistic species with simple morphologies. Further, these patterns resulted in a progressive decrease in species richness from the tropics to the poles.

the ectotherms, metabolic rates decline with temperature. Physiologically, the rate of reactions doubles for every 10 C increase in temperature. That means that four times as much food is required to support an organism in tropical waters as compared to subpolar waters. Thus, the rate of nutrient input that supports oligotrophic communities in warm waters might support mesotrophic or even eutrophic communities in cold waters. A minimal four-fold increase in food supply, associated with a drop in temperature of only few degrees during seasonal low-latitude upwelling, can shift a community from oligotrophic warm-water taxa to eutrophic cool-water taxa. Life-history strategy Organisms have been categorized by their life-history strategies, or their reproductive potential as related to competition and resource utilization (MacArthur and Wilson, 1967; Valentine, 1973). At one extreme are r-selected opportunists, able to rapidly increase their population densities usually by early maturation and faster reproduction (Hallock, 1985). Opportunists proliferate in resource-rich, low-stability regimes where their numbers fluctuate widely. At the other extreme are K-selected specialists, characterized by long individual life and low reproductive potential. The K-selected strategy is most advantageous in highly stable, typically oligotrophic environments where organisms compete by specialization and habitat partitioning. In between are a range of organisms adapted to mesotrophic environments, where tendencies towards r- or K-selected strategies vary relative to the two end members. In planktonic foraminifers, size is often directly related to reproductive potential; the faster a foraminifer matures and reproduces, the higher its reproductive potential (Hallock, 1985). Generally, small-sized foraminifers tend more towards the rselected end of the reproductive spectrum, whereas large-sized foraminifers tend towards the K-selected end (Caron and Homewood, 1983; Hallock et al., 1991). Opportunistic planktonic species are typically small, morphologically variable, widely distributed, and rapidly increase in abundance when nutrients become available. K-selected planktonic species may be larger, morphologically more complex and specialized, host algal symbionts, and are typically most diverse in oligotrophic waters where they compete by specialization and habitat partitioning. In terms of the sedimentological record and fossilization potential for the past, plankton diversity is high below oligotrophic and eutrophic environments because high proportions of such biotas produce calcareous and siliceous shells, respectively. While pelagic carbonate production is active in oligotrophic to mildly eutrophic waters (Hallock et al., 1991), under truly eutrophic conditions, short diatom-based food chains dominate at the expense of most planktonic foraminiferal and coccolithophorid taxa. In addition, high CO2 concentrations in subsurface waters in eutrophic regions severely limit carbonate preservation. Planktonic foraminiferal diversity, similar to that of other organisms, should be maximal in boundary regions domi-

Downloaded from specialpapers.gsapubs.org on March 18, 2011

Cretaceous paleoceanography: Evidence from planktonic foraminiferal evolution In terms of vertical stratification, another useful modern parameter, Sliter (1972), Hart (1980), Caron and Homewood (1983), Caron (1983), Hart and Ball (1986), and Leckie (1987, 1989) suggested that the hedbergellids and heterohelicids lived in surface waters and were the most ubiquitous forms on the shelves. The same authors proposed that the ornamented globular chamber-bearing forms lived slightly deeper in the water column, the double-keeled globotruncanids lived deeper yet, and the single-keeled forms inhabited the deepest parts of the spectrum. This hypothesis can only be partially tested through isotopic analysis and/or through depth transects from near-shore to open-marine environments. The hypothesis is poorly supported by stable isotopes whose values show little interspecies variation even during the Maastrichtian (DHondt and Arthur, 1995, 1996) when Cretaceous bioprovinciality was at its strongest (Davids, 1966). Still, the degree of vertical stratification in Cretaceous planktonic foraminifers is unresolved because isotopic values of Cretaceous planktonic foraminifers are too scarce to allow a reliable reconstruction of their habitats and latitudinal and vertical temperature gradients in the Cretaceous ocean were weak (Douglas and Savin, 1975; Boersma and Shackleton, 1981; Barrera and Huber, 1990). Further, most of the ad hoc depth transects completed to date are located in middle latitudes (Hart, 1980) where the presence/absence of ornamented forms cannot be related unambiguously to depth habitat but may be due to climatic conditions. The only data from shallow-water tropical regions is derived from the shelf edge of two Pacific guyots, Wodejebato and MIT (Erba et al., 1995a), and is limited to very narrow time spans. At Wodejebato, Heterohelix and Archaeoglobigerina were found consistently with rare specimens of Gansserina, Rugoglobigerina, and Globotruncana in very low-diversity Maastrichtian faunas deposited during times of more open-marine conditions. Very rare specimens of Globotruncanita, Pseudoguembelina, Pseudotextularia, and Globigerinelloides also were found occasionally. The more limited planktonic foraminiferal assemblages of late Aptian and late Albian age from MIT Guyot yielded rare specimens of hedbergellids and globigerinelloidids among which, however, H. trocoidea, G. ferreolensis, a doubtful G. algerianus, and one possible rotaliporid were identified (Erba et al., 1995a). These latter records, however, are too fragmentary to be fully useful. As a consequence, we derive information about the life strategy of Cretaceous planktonic foraminifers from their reported latitudinal distribution (Table 1). Although the record is far from complete, the recent campaigns of the Ocean Drilling Program (ODP) provided a discrete latitudinal coverage from the Antarctic margins to the northern high middle-latitudes (Hart, 1980; Ciesielski et al., 1988; Huber, 1990, 1991; Leckie, 1990; Haig, 1992; Premoli Silva, 1992; Wonders, 1992). Next we examine the high-latitude record to augment the wealth of information from low latitudes (Coccioni and Premoli Silva, 1994; Premoli Silva and Sliter, 1994, and references therein) as the basis for our worldwide Cretaceous paleoceanographic interpretations. Northern high latitudes

309

The oldest record outside the tropical Tethys, north or south, is from the North Sea area where Barremian, probably late Barremian, planktonic foraminiferal assemblages are composed of several hedbergellids and a few globigerinelloidids similar to assemblages from low latitudes (Ascoli, 1976; Banner and Desai, 1988; Coccioni and Premoli Silva, 1994; Banner et al., 1993). This similarity can be extended to early Aptian assemblages, whereas the late Aptian and most of the Albian assemblages were still composed of hedbergellids and globigerinelloidids and lacked the ticinellids north of 35 paleolatitude (Hart, 1976). In the latest Albian, assemblages from southern England, although still dominated by hedbergellids, yielded few or occasionally common, ornamented taxa such as Rotalipora appenninica, Planomalina buxtorfi, Praeglobotruncana delrioensis, and Costellagerina (Magniez-Jannin, 1981). During the Cenomanian, the northern limit for rotaliporids moved further north reaching the Danish coasts in the Rotalipora reicheli Zone (Hart, 1979). Thereafter, although ornamented taxa continued to be recorded from mid-high northern latitudes, the surviving species displayed short stratigraphic ranges and represented a minor component of the assemblages that was increasingly overwhelmed by hedbergellids, whiteinellids, and later the archaeoglobigerinids (Robaszynski et al., 1979; Caron, 1985). Among the marginotruncanids, M. marginata was the most abundant, had the broadest latitudinal range, and may have appeared earlier in higher latitudes than in the Tethys according to Hilbrecht et al. (1992). Other species recorded from high latitudes are M. renzi and M. pseudolinneiana associated with Dicarinella canaliculata in the Coniacian and Globotruncana linneiana in the Santonian, whereas the Dicarinella concavata group is recorded discontinuously (Bailey and Hart, 1979). Similar assemblages are recorded from the Great Valley sequence of northern California perhaps influenced by the southward flowing eastern boundary current (see Douglas, 1969). The most northerly record of planktonic foraminifers is from the Arctic slope of Alaska where rare specimens of Hedbergella and Heterohelix are present in the middle Turonian (Tappan, 1962). During the Campanian, ornamented planktonic foraminifers were greatly reduced in high-latitude assemblages (Berggren, 1962) and faunas consist of limited Heterohelix (Stenestad, 1969), Archaeoglobigerina, and Globigerinelloides (Caron, 1985). By the late Campanian through Maastrichtian, the assemblages were dominated by Rugoglobigerina along with a few Globigerinelloides and Heterohelix (Berggren, 1962). Ornamented species migrated as far north as southern Scandinavia twice during this interval (Berggren, 1962); Globotruncana linneiana and Rugotruncana subcircumnodifer in the late Campanian, and a much more diversified fauna during the late Maastrichtian Abathomphalus mayaroensis Zone. This latter fauna included Globotruncana arca, G. rosetta, G. mariei, Contusotruncana contusa, Globotruncanita stuarti, Rugoglobigerina macrocephala, Kuglerina rotundata, Racemiguembelina fructi-

Downloaded from specialpapers.gsapubs.org on March 18, 2011

Downloaded from specialpapers.gsapubs.org on March 18, 2011

Downloaded from specialpapers.gsapubs.org on March 18, 2011

312

I. Premoli Silva and W. V. Sliter

cosa, and Pseudotextularia elegans along with globotruncanellids and Abathomphalus. A similar assemblage is recorded on the Nova Scotian shelf and Orphan Knoll in the northwestern Atlantic. The assemblage includes the above species plus Gansserina gansseri, less common rugoglobigerinids, very common Heterohelix, and, among the hedbergellids, H. monmouthensis is well represented (Davids, 1966; our personal observation). Davids (1966) also noted the absence of Globotruncanella citae (=G. havanensis) north of 50 latitude. Southern high latitudes In the austral realm, the oldest planktonic foraminifers recorded are Caucasella hauterivica and Guembelitria sp. from Site 765 in the Argo Abyssal Plain off northwestern Australia that are probably early Aptian in age (Gradstein, Ludden et al., 1990). Upper Aptian sediments and planktonic faunas were recovered higher in the sequence at the same site, as well as nearby at Site 766 (Haig, 1992) and on Maud Rise (Leckie, 1990). Recurrent species are Hedbergella planispira and H. delrioensis whereas H. trocoidea, H. gorbachikae, Globigerinelloides ferreolensis, G. barri, H. infracretacea, and possible H. sigali are less common. Clavate forms are missing from this interval at all sites. A very depauperate fauna is recorded through most of the Albian, where the most consistent species are H. planispira and H. delrioensis, except from Exmouth Plateau where assemblages are limited to H. planispira (Wonders, 1992). As noted by Haig (1992), a uniform planktonic fauna consisting of common Planomalina, Rotalipora, Praeglobotruncana, Clavihedbergella, and Hedbergella occupied the latitudinal range between the Papuan Basin and Site 766 in the latest Albian. At the Exmouth Plateau, assemblages are still dominated by hedbergellids and contain Planomalina buxtorfi and Praeglobotruncana delrioensis but no rotaliporids (Wonders, 1992). In contrast, the coeval fauna on Naturaliste Plateau (Site 258) is much less diversified and consists of rare specimens of Hedbergella planispira, H. delrioensis, Clavihedbergella simplex, and Praeglobotruncana delrioensis (Herb, 1974) in the absence of Rotalipora and Planomalina. Of interest is the absence of true Ticinella at all sites.

In the Cauvery Basin at the tip of the Indian subcontinent, planktonic foraminiferal assemblages from questionable latest Aptian through most of the Albian contain diverse hedbergellids, Globigerinelloides bentonensis, G. caseyi, and rare Ticinella roberti (Venkatachalapathy and Ragothaman, 1995). In the latest Albian, however, the Rotalipora appenninica Zone assemblage also yielded R. balernaensis and Praeglobotruncana stephani (Venkatachalapathy and Ragothaman, 1995). Complete Tethyan faunas are recorded from the Ladack and Nepal areas, both of which are located at the northern edge of the Indian subcontinent. There, the biostratigraphic succession from the Rotalipora subticinensis Subzone through the R. appenninica Zone was identified (Premoli Silva et al., 1991). According to the reconstructions of Golonka et al. (1994) and Ricou (in Baumgartner et al., 1992) and Dercourt et al. (1993), the western Antarctic and northwestern Australian margins were located between 55 S and 6062 S latitude in the early Aptian whereas the Naturaliste Plateau lay some 10 southward of this belt and the Cauvery Basin was almost at the same latitude as Maud Rise at 64 S. In the late Albian, plate motion moved the Australian margin and India about 10 northward whereas Antarctica probably did not change latitudinally. Middle to late Cenomanian assemblages from Exmouth Plateau are more similar to those of the Tethys with rotaliporids and praeglobotruncanids (Wonders, 1992). Rotalipora appenninica is very rare but R. brotzeni, R. deeckei, and R. reicheli in the presence of R. cushmani are apparently more common than in the Tethys. Rotaliporids became extinct just prior to the Bonarelli equivalent and whiteinellids are common across the Cenomanian/Turonian boundary. Helvetoglobotruncana helvetica, Dicarinella hagni, and D. imbricata are present as well as the marginotruncanids, which evolved biostratigraphically as in the Tethys. Important components of the faunas are Falsotruncana maslakovae and Marginotruncana marianosi along with Hedbergella flandrini. All species of marginotruncanids are apparently present, including the large marginotruncanids, whereas the Dicarinella concavata group is poorly and discontinuously represented. Globotruncanita elevata seems to appear at the same level as in the tropics. Odd is the highest record of Marginotruncana sinuosa, M. undulata, and H. flandrini within the lower

Downloaded from specialpapers.gsapubs.org on March 18, 2011

Cretaceous paleoceanography: Evidence from planktonic foraminiferal evolution Campanian and well above the first occurrence of the nannofossil Aspidiscus parcus (=basal Campanian) and the last occurrence of D. asymetrica. Typically these species do not extend into the Campanian (see Caron, 1985; Premoli Silva and Sliter, 1994) and their occurrence, if not reworked, may represent an extension of their ranges in southern high latitudes. Depauperate faunas occur in most of the Campanian and through the lower half of the Maastrichtian on Exmouth Plateau. Globotruncana arca, Globigerinelloides, and Heterohelix (especially H. globulosa) are dominant and G. bulloides, C. fornicata, and G. linneiana are present in discontinuous pulses (Wonders, 1992). Rugoglobigerina rugosa, R. milamensis, Globotruncanella havanensis, Gublerina sp., and Planoglobulina carseyae first occur at the appropriate biostratigraphic levels. Late Maastrichtian faunas again exhibit strong Tethys affinities with C. contusa and A. mayaroensis among others, whereas Globotruncanita, Gansserina, and G. rosetta are poorly represented. Antarctic assemblages in the Campanian and Maastrichtian are dominated by Heterohelix and Globigerinelloides (Huber, 1990). Hedbergellids are common but never reach the abundance of the previous two groups. The hedbergellid maximum abundance is related to H. sliteri, which is endemic to the Antarctic region, whereas H. holmdelensis and H. monmouthensis often are poorly represented or even absent in the late Maastrichtian. The archaeoglobigerinids, frequently represented by endemic species, are common at Maud Rise but less common in subantarctic regions. Abathomphalus mayaroensis is present or even abundant along the Antarctic margin and apparently appears earlier there than in the tropics (Huber, 1992). Globotruncanellids are represented by G. petaloidea and, in minor amounts, by G. havanenis and G. pschadae. Among the large heterohelicids only Pseudotextularia elegans and Gublerina are recorded. Keeled forms are rare and mainly are represented by the genus Rugotruncana (Huber, 1992). CRETACEOUS PLANKTONIC FORAMINIFERAL LIFE STRATEGIES Comparing the known latitudinal and vertical distributions of Cretaceous planktonic foraminifers to their morphologic characteristics provides a means to interpret their life strategies. As with modern planktonic foraminifers, we believe that the size of the Cretaceous taxa likewise was related to resource utilization and reproductive potential. Similarly, we employ the terms rselected and K-selected to describe the reproductive strategy of the Cretaceous species while recognizing there is a continuum between these end members. The life history strategy for a given Cretaceous species in this continuum was related to a particular environment at a given time. To aid in this comparison we recognize the following 14 recurrent taxonomic clusters ranked according to progressively more complex gross test morphology, ornamentation, and wall-surface porosity: 1. Simple low-trochospiral morphotypes with globular chambers, which include the thin- and smooth-walled hedbergellids.

313

2. Low to medium-high trochospiral morphotypes mainly with subglobular chambers and possessing marked, more or less organized rugosities on the surface; these include Costellagerina, Whiteinella, Archaeoglobigerina, Rugoglobigerina, Kuglerina, and Trinitella. 3. Low trochospiral morphotypes with subglobular chambers, secondary apertures, and a tendency to thicken the walls of the inner whorls; the ticinellids. 4. Low trochospiral morphotypes with subacute to acute peripheral margins lacking true keels; includes Praeglobotruncana and Globotruncanella. 5. Trochospiral morphotypes, mainly planoconvex, possessing juvenile subglobular chambers and hemispherical chambers in the adult with or without marginal keel(s); these include Dicarinella, Helvetoglobotruncana, Gansserina, and possibly Rugotruncana. 6. Single- to double-keeled, mostly trochospiral morphotypes with acute and/or truncated margins and raised sutures on both spiral and umbilical sides; these include Rotalipora, Marginotruncana, Globotruncana, Contusotruncana, and Globotruncanita. We include the ornamented planispiral genus Planomalina in this group. 7. Double-keeled, trochospiral morphotypes with truncated margins and raised sutures only on the spiral side; includes Falsotruncana and Abathomphalus. 8. Planispiral morphotypes, mainly small-sized, with compressed to inflated chambers; Globigerinelloides. 9. Thin, microperforate biserial heterohelicids with a smooth to finely striate surface; includes Heterohelix and Laeviheterohelix. 10. Biserial heterohelicids with supplementary apertures; Pseudoguembelina. 11. Flaring heterohelicids with more than two chambers per row; Ventilabrella, Planoglobulina, and Gublerina. 12. Medium-sized heterohelicids with chambers arranged from biserial to annular and a surface from smooth to costate; includes Pseudotextularia and Racemiguembelina. 13. Large, thick-walled plurigeneric cluster including Hedbergella trocoidea, Ticinella roberti, and Biticinella breggiensis, which display some features in common with groups 2 and 3 but lack the marked surface ornamentation, and Globigerinelloides algerianus and G. barri, which are planispiral like group 8. 14. Clavate to tubulospinous-chambered cluster; includes Clavihedbergella, Hedbergella rhinoceros, Leupoldina, Schackoina, Plummerita, and Radotruncana. These 14 clusters are further lumped into 10 discrete groups (A to J) with apparently similar environmental behavior. These groups are based on latitudinal distribution, from cosmopolitan to progressively more endemic. (A) The simplest morphologies in both trochospiral and biserial clusters (1 and 9) are the most cosmopolitan and opportunistic. (B) Small Globigerinelloides (cluster 8) behave similarly to the simplest group but are slightly less tolerant of variable conditions. (C) Praeglobotruncana and Globotruncanella (cluster 4) are the most cosmopolitan taxa among the more complex clusters. (D) The round chambered

Downloaded from specialpapers.gsapubs.org on March 18, 2011

314

I. Premoli Silva and W. V. Sliter they are rare and very small in size, then gradually increase both in size and abundance, reaching their maximum diversification in the Aptian Leupoldina cabri and Globigerinelloides ferreolensis Zones. During the late Aptian, the assemblages became enriched by the first r/K intermediate forms such as the large-sized, thickwalled Globigerinelloides algerianus, G. barri, Hedbergella trocoidea, Ticinella roberti, and T. bejaouaensis. As these intermediate forms disappear in the late Aptian, the early to middle Albian planktonic foraminiferal assemblages were once again dominated initially by small-sized opportunists, but they rapidly acquired new r/K intermediate forms represented by the ticinellids, first of small size then progressively of increasing size and complexity (see Brhret et al., 1986). Morphologic diversity similar to that of the late Aptian resumed close to the base of the Biticinella breggiensis Zone. This is followed by the continued diversification and size increase of the r/K intermediate group. The late Albian marks the appearance of the first K-selected strategists, the rotaliporids, which rapidly diversified. Diversification, however, also continued among the r-selected opportunist and r/K intermediate groups with the appearance of thin, biserial Heterohelix and praeglobotruncanids, respectively. The maximum diversity in this period was reached in the lower part of the Rotalipora appenninica Zone where the stratigraphic distribution of new taxa overlaps with older r/K intermediate forms, such as the ticinellids. After a period during which the planktonic foraminiferal community remained rather stable (early to early late Cenomanian), a new diversification within the r/K intermediate group in the late Cenomanian announced the organic carbon-rich (Corg-rich) Bonarelli Event. The Bonarelli Event (latest Cenomanian) is characterized by the dominance of the r/K intermediate whiteinellids associated with abundant r-selected opportunists and much rarer r/K dicarinellids in the absence of the K-selected strategists. K-strategists reappear in the middle Turonian, represented by the genus Marginotruncana, which rapidly diversifies. These species are joined in the Coniacian to Santonian by the first occurrence of the globotruncanids, contusotruncanids, and globotruncanitids, contributing to the high diversity in the Late Cretaceous. The highest diversification within the K-strategists is reached in the late Campanian to early Maastrichtian, followed by a slow decline in diversity toward the end of the Maastrichtian. The decreasing diversity, however, is not paralleled by a decrease in abundance of the K-strategists, except just prior to the K/T boundary. An increasing diversification within the r/K intermediate group accompanies that of the K-strategists with the appearance of new genera and species. However, their abundance in the assemblages commonly is modest except at very high latitudes where they may dominate the faunas, such as the archaeoglobigerinids in the Maastrichtian from the Antarctic margin (Huber, 1990, 1992). Moreover, representatives of the r/K intermediate group became progressively more important components of the assemblages when the K-selected strategists decreased in diversity just prior to the K/T boundary. It is also worth mentioning

pustulose forms (equated to cluster 2); among them Whiteinella and Archaeoglobigerina tend to display high dominance in their respective assemblages, suggesting higher tolerance to eutrophic conditions and a broader latitudinal distribution with respect to the other genera of the cluster. (E) Ticinellids (cluster 3), planoconvex forms (cluster 5), and the various species of cluster 13 display similar behavior in their respective time intervals; they are mainly missing at high latitudes. (F) Rare complex heterohelicids of clusters 10 and 12 are occasionally present in high-latitude assemblages. (G) Among the flaring heterohelicids (cluster 11), Gublerina is the only form recorded, although rarely, at high latitudes. (H) Falsotruncana and especially Abathomphalus (cluster 7) may be important components of the middle- to highlatitude assemblages in their respective time intervals and are much rarer at low latitudes except for the higher abundance of Falsotruncana in marginal seas such as in Tunisia (Caron, 1981). (I) The keeled cluster (6) includes less tolerant forms that dominate low- to middle-latitude assemblages with high diversity. Among them, Marginotruncana marginata, M. coronata, M. pseudolinneiana, Globotruncana bulloides, and G. linneiana display a slightly broader latitudinal range than the other forms in this cluster. (J) Representatives of the clavate cluster (14) show a very discontinuous stratigraphic range with a distinct preference for low to middle latitudes. In terms of life strategy (Fig. 4, Table 2), the simplest morphologic groups (A and B above) are opportunists, or r-strategists, with high reproductive potential and inhabiting more nutrient-rich waters close to the eutrophic part of the resource spectrum. The number of species in this group is small throughout the entire Cretaceous. In contrast, the complex, large-sized keeled morphotypes (I above), are the most specialized and least tolerant to environmental variation, or K-strategists, that inhabit low-nutrient waters close to the oligotrophic part of the spectrum and compete by habitat partitioning. This group is highly diversified. In between these reproductive end members are the r/K intermediate morphotypes that display a range of trophic strategies. In Table 2, we have attempted to place members along the continuum as more r- or K-selected. The number of species in each genus in the intermediate group is small (see Fig. 4) but the group is highly diversified overall, although less diversified than the K-strategists. In particular, the high dominance behavior of Whiteinella and Archaeoglobigerina suggests that these forms had a high reproductive potential that seems to characterize upwelling regions. Finally, the clavate group in analogy with Eocene hantkeninids may have inhabited oxygen-depleted surface to near-surface waters (Boersma et al., 1987). Life strategies with time As shown in Figure 4 , the abundance of the r-selected opportunists, r/K intermediate group, and K-selected strategists changed through time. In the Early Cretaceous, planktonic foraminiferal assemblages are composed only of r-selected opportunistic hedbergellids and globigerinelloidids: at the beginning

Downloaded from specialpapers.gsapubs.org on March 18, 2011

Cretaceous paleoceanography: Evidence from planktonic foraminiferal evolution

315

Figure 4. Stratigraphic distribution of Cretaceous planktonic foraminifers grouped according to their inferred life strategy (data from Fig. 2) plotted against the main ecological changes (chronology as in Fig. 1).

that the r-selected opportunists, the most consistent component of the Cretaceous planktonic faunas, are hedbergellid-dominated in the Early and early Late Cretaceous, but later in the Late Cretaceous they become heterohelicid-dominated in both diversity and abundance. Representatives of the clavate group display a random stratigraphic distribution (Fig. 4) with schackoinids the most consistent forms in low latitudes (Table 2). Most of the time they are represented by rare specimens of a single species except in two intervals; in the Barremian to the early late Aptian, where they display the highest diversity, and in the late Campanian. Peculiarly, the Barremian to Aptian interval is characterized by the

widespread occurrence of several Corg-rich layers that culminated in the Selli horizon with a Corg content greater than 8% total weight (Weissert, 1989), but no Corg-rich layers are recorded in the late Campanian. As a consequence, the paleoceanographic significance of the clavate group is still ambiguous. Additionally, we note the following faunal characteristics: a. Hedbergellids are confirmed as shallow-water dwelling as well as the most tolerant of the trochospiral planktonics with several species described from high latitudes (Douglas and Rankin, 1969; Herb, 1974; Leckie, 1990; Venkatachalapathy and Ragothaman, 1995). b. Narrow heterohelicids are more tolerant than flaring types.

Downloaded from specialpapers.gsapubs.org on March 18, 2011

316

I. Premoli Silva and W. V. Sliter

The former are ubiquitous and often abundant, suggesting a life strategy close to that of the r-selected hedbergellids. c. Among more complex heterohelicids, Pseudoguembelina, Pseudotextularia, and Gublerina migrated into higher latitudes toward Antarctica. The data, however, are far from complete as heterohelicids often were overlooked in previous studies. d. Large and small forms with globular chambers with vari-

ous ornamentation show a life strategy more tolerant to environmental changes. Included here are species of Hedbergella, Ticinella, Costellagerina, Whiteinella, Helvetoglobotruncana, Archaeoglobigerina, and Gansserina. e. Among the ticinellids, T. primula is more tolerant than T. praeticinensis or T. roberti. f. The presence of Gansserina and Archaeoglobigerina on

Downloaded from specialpapers.gsapubs.org on March 18, 2011

Cretaceous paleoceanography: Evidence from planktonic foraminiferal evolution Pacific guyots (Erba et al., 1995a) and in the subantarctic region (Huber, 1991), and the latter genus from the Antarctic margin, suggests that both groups preferred a shallower-water habitat and are environmentally tolerant. This interpretation is supported by stable isotope values from Archaeoglobigerina (Barrera and Huber, 1990). g. The abundance and earlier appearance of Abathomphalus mayaroensis in middle to high southern latitudes suggests a preferred extratropical habitat for this r/K intermediate taxon (Huber, 1992; Huber et al., 1995). h. Marginotruncana marginata and Globotruncana bulloides are the most tolerant of the K-strategists to changes in the environment. CRETACEOUS PALEOCEANOGRAPHIC CHANGES WITH TIME The evolutionary changes that we observe in Cretaceous planktonic foraminifers provide a means to interpret the history of paleoceanographic changes. In the Late Jurassic to Early Cretaceous, the surface of the world ocean can be described as a rather eutrophic paleoenvironment inhabited by r-selected opportunists even though the record is limited mainly to rare low-latitude sites. Globuligerinids dominated until the first diversification of the Cretaceous planktonic foraminifers in the early Valanginian (Fig 5). This diversification coincides in general with several events in other pelagic fossil groups as well as with geochemical and sedimentological events. For example, hedbergellids first occured as the calpionellids drastically decreased in abundance shortly before their extinction. At almost the same level, nannoconids, the main constituent of the Early Cretaceous low-latitude pelagic sediments (Erba and Quadrio, 1987), decrease in abundance whereas coccolithophorids diversify, boreal dinoflagellates migrate to low latitudes (Leereveld, 1995), and diatoms become somewhat better preserved (Erba and Quadrio, 1987). These events portend changes in the water column derived from the increased flow of horizontal and vertical oceanic currents. The biologic events slightly precede a positive shift in 13C that Weissert (1989) correlated to an increase in river discharge during a warm, humid, climatic regime. These conditions contributed to the increased preservation of organic matter. Correspondingly, the isotopic shift coincides with the first deposition of Corg-rich black shales intercalated with Cretaceous pelagic limestone. The increased seasonality from dry to more humid climatic conditions with increased fresh-water runoff likely increased the nutrient supply in surface waters, creating more eutrophic conditions suitable for diversification of the hedbergellids. The decrease in nannoconids at this time supports this interpretation. Erba (1994) suggested the nannoconids are analogous to extant Florisphaera, a nannofossil that proliferates when the deep chlorophyll maximum (DCM) lies in the lower part of the photic zone. On the other hand, shoaling of the DCM favors the proliferation of coccolithophorids in the upper part of the photic zone. If Erbas interpretation is correct, the hedbergellids may

317

have appeared when nutrients were concentrated in near-surface waters concomitant with the increased river discharge hypothesized by Weissert (1989; see also Lini et al., 1992) and upwelling in the upper water column. After their first diversification, planktonic foraminifers underwent a prolonged period of evolutionary stasis in the late Valanginian to earliest Barremian. At the same time, their stratigraphic distribution and abundance fluctuated rhythmically (Coccioni et al., 1992) in accordance with alternating limestone/marl bedding. These lithologic rhythms have been related to climatic changes induced by orbitally produced Milankovitch cycles (Herbert and Fischer, 1986). The existence at that time of discrete seasonality is supported by provincialism that affected calcareous nannofossils, ammonites, and belemnites in northern Europe (Mutterlose, 1991). Fluctuations in seasonality apparently accelerated and strengthened in the early Barremian inducing more contrasting nutrient levels in surface waters. New niches were developed for the gradual diversification of opportunists and the first occurrence of clavate forms represented by Clavihedbergella and Leupoldina. Planktonic foraminifers became abundant for the first time in pelagic carbonates and increased in size. The first occurrence of benthic genera such as Gavelinella, that characterize middleslope water depths later in the Cretaceous, suggests the development of an intermediate water layer (Sliter, 1980). Still, environmental conditions maintained the deep position of the DCM for nannoconid proliferation (Erba, 1994). The new niches for opportunists apparently extended into the northern high latitudes (Banner et al., 1993) and, although we lack control from the southern high latitudes, the planktonic fauna apparently was cosmopolitan. In the late Barremian, planktonic foraminifera continued to diversify but rhythmic fluctuations in abundance increased in intensity (Coccioni et al., 1992). Paleoenvironmental conditions were alternately more to less suitable for the globigerinelloidids as well, ranging from less eutrophic during their presence to more eutrophic during their absence. Such environmental fluctuations doubtless were a widespread phenomena and not confined to low latitudes, however; the spotty record from northern high latitudes does not provide the answer. The fact that benthic foraminifers increased in diversity and black-shale layers became more common in the same interval may indicate that the changing paleoenvironmental conditions affected the entire water column. The first turnover in planktonic foraminifers is associated with the Corg-rich deposition at the Selli Level (Figs. 3, 5). The event was preceded by a diversification of hedbergellids associated with the temporary disappearance of nannoconids (Erba, 1994) and a major turnover in radiolarians (Erbacher, 1994). Closely following or perhaps still associated with the waning phase of the increased burial of organic matter of marine origin and the second major Cretaceous positive shift of 13C, there is a temporary decrease of hedbergellids and globigerinelloidids coeval with a diversification within the clavate group. Paleoenvironmental conditions surrounding the Selli Event were proba-

Downloaded from specialpapers.gsapubs.org on March 18, 2011

Downloaded from specialpapers.gsapubs.org on March 18, 2011

Figure 5 (on this and previous page). Synopsis of the paleoenvironmetal changes through the Cretaceous derived from the evolution of planktonic foraminifers compared to events in other fossil groups and physico-chemical parameters (chronology as on Fig. 1 supplemented by calcareous nannofossil data). Vertical crosshatched bar in column for Planktonic foraminiferal evolutionary change shows the threefold evolutionary pattern. Main references: Arthur et al. (1990), Erba (1994), Erbacher (1994), Herbert and Fischer (1986), Herbert et al. (1995), Leereveld (1995), Lini et al. (1992), Roth and Krumbach (1986), Weissert (1989), Weissert and Brhret (1991), and Weissert and Lini (1991).

Downloaded from specialpapers.gsapubs.org on March 18, 2011

320

I. Premoli Silva and W. V. Sliter nensis Zone heralded for the first time during the Cretaceous the development of a pronounced stratification of the mixed layer above a well-defined thermocline. Paleoenvironmental conditions shifted to more oligotrophic even if still within the mesotrophic range as indicated by the as yet abundant r-opportunists. Climatedriven rhythmic fluctuations within Milankovitch frequencies dominated during this period of subtle environmental changes (Herbert and Fischer, 1986; Premoli Silva et al., 1989b; Tornaghi et al., 1989; Erba, 1992; Herbert et al., 1995). The onset of a stronger thermocline requires an increase in equator-to-pole temperature gradients. The expected increase in temperature gradient is supported by the occurrence of less diversified planktonic foraminiferal assemblages on the Antarctic margin (Herb, 1974), indicating that a weak bioprovince perhaps was differentiated at the highest southern latitudes (>60 S) by the late Albian (see Baumgartner et al., 1992). Diversification of planktonic foraminifers is paralleled by diversification within radiolarians during most of the Albian; moreover, the foraminiferal turnover and extinction event in the upper part of the Rotalipora appenninica Zone also is registered within the radiolarians (Erbacher, 1994). The widespread hiatus at this level (Schlanger, 1986) and the apparent coeval demise of several carbonate platforms in the Atlantic and Pacific Oceans (e.g., Hallock and Schlager, 1986; Winterer et al., 1995; Haggerty et al., 1995) appear to be associated with a change in paleotemperature (Erbacher, 1994; Sliter, 1995). This late Albian event was followed by a period of evolutionary stasis that spanned the early Cenomanian (Fig. 5). This stasis was associated with the expansion of warm niches toward higher latitudes. In the Northern Hemisphere, low-latitude planktonic foraminiferal assemblages with K-selected strategists first occurred on the English margin during the Cenomanian transgression. These faunas show a strong similarity to austral temperate faunas from Exmouth Plateau (Wonders, 1992) that by the Cenomanian had been transported northward through oceanic plate motion. During most of the Cenomanian, however, the southern high-latitude bioprovince had become more discrete and was characterized only by rare K-selected strategists at the latitude of the Falkland Islands and Tierra del Fuego (Sliter, 1976; Krasheninnikov and Basov, 1983), much farther north than during the late Albian. In the late Cenomanian, diversification among planktonic foraminifers occurred primarily within r/K intermediate forms and to a minor extent within r-selected opportunists, whereas Kselected strategists decreased in diversity. These trends indicate a further change toward less stable and more eutrophic paleoenvironmental conditions that culminated in the Corg-rich Bonarelli Event. At that time the thermocline was once again disrupted and surface to subsurface waters became very unstable and controlled by an upwelling regime. The Bonarelli Event, the worldwide OAE 2 (see Arthur et al., 1990), coincides with the highest accumulation of organic matter of marine origin in pelagic sediments during the Cretaceous (>23% total weight; Arthur and Premoli Silva, 1982), a

bly quite peculiar, with increased upwelling and the probable expansion of oxygen-depleted waters (Arthur et al., 1990). Still, these conditions did not prevent diversification within the hedbergellids and globigerinelloidids. Several interpretations regarding the significance of the Selli Event have been put forward. Larson (1991) related it to the acme of volcanic activity, or superplume, which disrupted the previous steady state leading to the temporary nannoconid crisis (see Erba, 1994). Erbacher (1994) spoke of an increase in productivity due to enhanced nutrient input into surface water from flooded coastal areas during a rise in sea level. This increased productivity caused an expansion of the oxygen minimum zone. That these events reached to oceanic surface waters is attested to by the turnover in planktonic foraminiferal opportunists. Following these events, planktonic foraminifers rapidly increased in size and diversity. Niche-partitioning of the mixed layer is evident by the middle of the late Aptian, as indicated by the first occurrence of large, r/K intermediate hedbergellids and thick-walled globigerinelloidids. Nannoconids registered a new acme (N. truittii Acme, see Erba, 1994), 13C values became more positive (Weissert, 1989), and radiolarians underwent gradual extinction (Erbacher, 1994). The appearance of the intermediate planktonic foraminifers signals the onset or strengthening of a weak thermocline. For the first time, surface waters were characterized by less eutrophic, probably oligotrophic to mesotrophic, paleoenvironmental conditions. At the same time, the occurrence of localized black shales such as those of the Calera Limestone, originally located in the ancestral Pacific Ocean basin (Sliter, 1989), attest to continued, if brief, fluctuations in the strength of the thermocline. The post-Selli paleoenvironmental conditions expanded toward high southern latitudes but apparently not into northern high latitudes. Northern assemblages consist only of opportunists, perhaps reflecting the shallow-water depths of the reported sites as the nannoconid acme is also recorded in northern Europe (Mutterlose, 1989; Erba, 1994). This episode of diversification was interrupted before the close of the Aptian when the weak thermocline was completely disrupted, provoking the extinction and temporary disappearance of r/K intermediate forms. The early Albian planktonic foraminiferal assemblages once again were composed of only r-selected opportunists, suggesting the return of eutrophic conditions in a likely well-developed upwelling regime. These more eutrophic conditions were accompanied by a new Corg-rich accumulation event (OAE 1b; Arthur et al., 1990; Erba, 1994), strong carbonate dissolution (Brhret et al., 1986; Premoli Silva et al., 1989a; Erba, 1992), and a turnover among radiolarians (Erbacher, 1994). Planktonic foraminifers did not diversify during this period. As the upwelling regime decreased, niche-partitioning within the mixed layer gradually resumed, first accommodating the r/K intermediate group and then returning to a late Aptian state by the Ticinella praeticinensis Subzone. Contrary to the late Aptian, diversification continued into the late Albian and the occurrence of the first K-selected strategists in the Rotalipora tici-

Downloaded from specialpapers.gsapubs.org on March 18, 2011

Cretaceous paleoceanography: Evidence from planktonic foraminiferal evolution major turnover within radiolarians (e.g., Marcucci Passerini et al., 1991; Thurow et al., 1992; Erbacher, 1994), mollusks (Elder, 1989), and calcareous nannofossils (Erba et al., 1995b), and a major positive shift of 13C. Several authors have associated this event with a pulse of very high primary productivity induced by an exceptional increase in nutrients in surface and subsurface waters (Herbin et al., 1986; Arthur et al., 1987; Coccioni et al., 1991; Bellanca et al., 1996). The Bonarelli Event is preceded by precursor events such as numerous black-shale layers and the increasing abundance of Heterohelix and r/K intermediate forms (whiteinellids and dicarinellids) paralleled by the gradual extinction of the K-selected rotaliporids. In terms of diversity, a short period of stasis followed the Bonarelli Event (Fig. 5). This interval is unusual, however, as the gradually decreased dominance of the whiteinellids is balanced by the increased dominance of the praeglobotruncanids, dicarinellids, heterohelicids, and hedbergellids. The appearance of Helvetoglobotruncana helvetica, a single-keeled descendant of the rounded-shaped whiteinellids, indicates that the upper water column slowly restratified. The thermocline again was well defined by the middle Turonian, when the K-selected marginotruncanids first occurred, and remained a permanent feature of the Late Cretaceous ocean until the end of the era. Shortly after the increased stratification of the upper water column, K-selected strategists became highly diversified and dominated the large size-fraction of the assemblages, while the r-selected opportunists were still abundant in fine fraction. The r/K intermediate group underwent a marked turnover in the late Turonian with the extinction of the praeglobotruncanids and some of the dicarinellids inherited from the latest Cenomanian. As a result, despite the appearance of the archaeoglobigerinids, the intermediate group overall became a minor component of the assemblages (Fig. 4). This situation continued with little change for more than 4 m.y. From the Bonarelli Event through the Coniacian, tropical paleoenvironmental conditions expanded over a large latitudinal band extending to the subantarctic region and provinciality decreased. Only a few K-strategists seemed better adapted to high latitudes than to low latitudes such as Marginotruncana marianosi, Falsotruncana, and Hedbergella flandrini, all of which were more abundant outside the tropical belt to at least the southern mid-high latitudes. However, the spotty record of true highlatitude oceanic sites for this time period prevents understanding the complete latitudinal provincial gradient. For example, the presence of two opportunistic species from Turonian strata on the Arctic margin (Tappan, 1962) likely reflects a depauperate fauna associated with coastal shallow-water environments. The largest turnover of planktonic foraminifers in the Cretaceous occurred in the Santonian Dicarinella asymetrica Zone and affected all trophic groups, (Fig. 5). Both originations and extinctions exemplify the interval (Fig. 3), however, the origination of new genera and species outnumbered extinctions (Fig. 2). Diversity is high and new forms include new morphologies such as those of the compressed heterohelicid Laeviheterohelix, the

321

flaring heterohelicids Sigalia and Ventilabrella, inflated globigerinelloidids, and the clavate Eohastigerinella. The interval represents the transition between K-selected faunas dominated by marginotruncanids to one dominated by globotruncanids and globotruncanitids (Wonders, 1980). The interval further represents the ecotone between two major ocean types; the earlier Greenhouse ocean with variable sediments consisting of multicolored carbonate typical of redox cycles, chert, and black shale and with weak bioprovinces, and the later modern ocean dominated by more uniform carbonate and with well-defined bioprovinces. Accompanying these changes in the D. asymetrica Zone is a major turnover in calcareous nannofossils, the last occurrence of widespread black-shale deposition (Zimmerman et al., 1987), a strong episode of carbonate dissolution close to the top of the zone, a pulse of volcanic activity, and the final opening of a deepwater connection between the North and South Atlantic Oceans (Kennett, 1982). The latter event may have been key to understanding the scope of paleoceanographic changes. Opening of full oceanic communication through a wider South Atlantic altered global deep-water circulation as noted by Wonders (1980) and likely precipitated changes in ocean chemistry, marine climate, and water structure. The unusual combination of both major originations and extinctions in all trophic groups indicates that changes occurred throughout the water column and supports the interpretation of an altered deep-water circulation. In the Campanian and Maastrichtian, tropical planktonic foraminiferal assemblages are dominated by highly diversified K-selected strategists that now include abundant globotruncanids, fewer globotruncanitids and Contusotruncana, and, among the opportunists, thin heterohelicids. In contrast, the formerly important hedbergellids became very rare. The r/K intermediate group increased considerably in diversity with the appearance of new morphotypes in both the complex heterohelicids and trochospiral forms. This signifies that the number of niches within the mixed layer increased measurably with respect to previous times to accommodate the more than 60 species recognized in low latitudes. This extreme partitioning implies that tropical surface waters were characterized by oligotrophic conditions as they are today. Moreover, as in the modern ocean, niches for K-selected strategists were progressively eliminated toward high latitudes. As expected, species diversity also decreased with increased latitude, allowing the identification of discrete bioprovinces that extended to the Antarctic margin (see also Davids, 1966). The onset of the bioprovinces occurred in the early Campanian during a period of taxonomic stasis and became more pronounced as the diversity of K-selected strategists increased in the late Campanian. Near the base of the Maastrichtian, the decrease in diversity within K-selected strategists and the opposite within the r/K intermediate group indicate a shift to less oligotrophic conditions accompanied by a progressive weakening of the thermocline. This trend strongly accelerated near the end of the Maastrichtian, except for a brief episode in the early Abathomphalus mayaroen-

Downloaded from specialpapers.gsapubs.org on March 18, 2011

322

I. Premoli Silva and W. V. Sliter canids range latitudinally from northern Europe to the subantarctic regions and extratropical provinces are defined by species and not genera. Provincialism again became well established in the early Campanian through the Maastrichtian (Huber and Watkins, 1992). The number of species characterizing the polar province, however, was much higher than in the modern ocean and more similar to faunal diversity in the Paleogene (Boersma and Premoli Silva, 1991). This means that polar water temperatures were well above present-day temperatures and thus supports the mean temperature of 10 C derived from isotopic analyses of Maastrichtian planktonic foraminifers from Antarctica (DHondt and Arthur, 1995). Moreover, even though Maastrichtian bioprovinces were well established, warming pulses are indicated by the migration to higher latitudes of more specialized taxa, either r/K intermediate forms or K-selected strategists, such as occurred during the late Maastrichtian 66.7-Ma episode recorded in Antarctica (Huber and Watkins, 1992) and northern Europe (Caron, 1985). 4. Throughout the Cretaceous, the onset of bioprovinces is better documented in the southern ocean as a consequence of the more widespread oceanic conditions at southern latitudes, whereas shallower seas and seaways were predominant at high latitudes in the Northern Hemisphere. The Arctic Ocean experienced stronger continentality than the Antarctic because it was surrounded by exposed lands. Consequently, Arctic planktonic faunas consisted of a small biomass composed mainly of hedbergellids. 5. Five major episodes interrupted the general evolutionary trend during the Cretaceous; the Selli Event, the Aptian/Albian boundary event, the Bonarelli Event, the Santonian event, and the extinction event at the end of the Cretaceous. Four of these events (the Selli Event, the Aptian/Albian boundary event, the Bonarelli Event, and the extinction event at the end of the Cretaceous), display physical as well as biological similarities but each has a unique geologic character as follows: a. The acme of all four events is preceded by precursor changes such as rhythmic fluctuations in carbonate content and calcareous plankton composition and abundance at orbital frequencies, and the progressive elimination of niches accommodating the least tolerant planktonic foraminifers. As a result, planktonic foraminiferal communities shifted from K-specialistto r/K intermediate- or r-opportunist dominated during the events. This last feature is absent prior to and across the Selli Event as assemblages at this evolutionary level still consisted solely of r-selected opportunists and the thermocline likely was absent or weakly developed. b. The first three events, the Selli, Aptian/Albian, and Bonarelli, are associated with the accumulation or enhanced preservation of organic matter of marine origin (OAE1a, OAE1b, and OAE2, respectively), a positive shift of 13C, and a turnover in radiolarians. In each case, the main black-shale bed of the event is underlain by black-shale layers that become progressively thicker and more frequent approaching the event.

sis Zone (at 66.7 Ma according to Huber and Watkins, 1992), when a few specimens of rugotruncanids and Globotruncana bulloides were recorded on the Antarctic margin (Huber, 1992), and Contusotruncana contusa migrated into the North Sea (Caron, 1985). At the close of the Cretaceous, the thermocline was totally disrupted and all niches disappeared. Planktonic foraminifers that survived this drastic event include the opportunistic Guembelitria, which flooded the ocean surface at the K/T boundary (i.e., Delacotte et al., 1985; Kroon and Nederbragt, 1990), very rare small hedbergellids (Keller, 1993; Olsson et al., 1992), and an ancestral form of the Paleocene chiloguembelinids (Huber and Boersma, 1994). This sparse planktonic fauna suggests that surface waters again became eutrophic. Campanian-Maastrichtian bioprovincialism is supported by isotope values that indicate the presence of a distinct temperature gradient from the tropics to poles. Barrera and Huber (1990) and DHondt and Arthur (1995, 1996) estimated a mean temperature of 10 C for Antarctic near-surface waters in the Maastrichtian. This warmer than modern temperature may account for the diverse high-latitude planktonic foraminiferal assemblages (Huber, 1990) that are more diverse than assemblages in the modern polar bioprovince (B, 1977) and similar to those of the Paleogene (Boersma and Premoli Silva, 1991). CONCLUSIONS Our interpretation of Cretaceous paleoceanography is based primarily on the evolutionary development (morphologic complexity, diversity, life strategy) and spatial distribution of planktonic foraminifers. Although we acknowledge the limitations placed on our interpretations by the scarcity of high-latitude data, we believe our analysis yields the following paleoceanographic and paleontologic conclusions: 1. Planktonic foraminifers apparently began to diversify when the DCM moved higher in the photic zone and surface waters became sufficiently rich in nutrients to support a phytoplankton prey suitable for planktonic foraminifers. 2. After the first diversification in the early Valanginian, planktonic foraminifers in general increased in diversity, size, and morphological complexity throughout the Cretaceous. The overall trend was related strictly to the availability of an increasing spectrum of environmental niches within the mixed layer as the thermocline became better defined and marine environments progressed from eutrophic to more oligotrophic conditions. 3. The development of a definitive thermocline and nichepartitioning was tied to increasing temperature gradients from the equator to the poles. As the diversity of the K-selected strategists increased, latitudinal bioprovinces became increasingly well differentiated. Although a weak provincialism is discerned back to the late Aptian, discrete bioprovinces characterized by the presence and absence of genera are first recognized during the latest Albian Rotalipora appenninica Zone, especially in the southern ocean. The bioprovincial gradient then lowers from the Cenomanian through Coniacian as H. helvetica and the marginotrun-

Downloaded from specialpapers.gsapubs.org on March 18, 2011

Cretaceous paleoceanography: Evidence from planktonic foraminiferal evolution c. The Selli, Aptian/Albian, and Bonarelli Events have been consistently associated with increased upwelling which created very unstable conditions in the upper water column and disrupted the thermocline, when present. As mentioned above, the changes in planktonic foraminiferal trophic structure and appearance of cyclic sedimentation prior to the K/T extinction event are very similar to those associated with the other three events. Similarly, we can conclude that the crisis at the K/T boundary also was due to an upwelling episode which had more drastic effects on the biomass than in previous times because the community structure, niche-partitioning, and the overall environmental conditions were much more specialized and extreme. In that context, the effects were catastrophic. This interpretation, however, is apparently contradicted by stable isotope data that indicate a drop in productivity (=negative 13C shift) at the boundary. The fifth or Santonian Dicarinella asymetrica Zone event separates the Greenhouse ocean from an ocean that resembles the modern oceans in the character of the bioprovinces, the trophic structure of planktonic foraminifers, and the type of marine sediments. This zone contains the largest turnover in Cretaceous planktonic foraminifers at all trophic levels. The magnitude of the changes suggests a modification of deep- and intermediate-water currents likely related to the onset of longitudinal deep-water circulation in the Atlantic Ocean. 6. Finally, although the overall Cretaceous climate was warmer and more uniform than today, Cretaceous paleoceanographic evolution occurred because of a progressive increase in seasonal climatic contrast that, at least, was partially associated with an increase in latitudinal temperature gradients needed to develop bioprovinces. ACKNOWLEDGMENTS We gratefully acknowledge the careful reviews by John Barron, Bill Berggren, Michele Caron, and Will Elder. In addition, we thank the editors for inviting us to participate in the volume and for their attention to detail. This paper was supported by a Ministry of University and Scientific Research and Technology grant, MURST 40%, to IPS.
Clavihedbergella simplex (Morrow) Contusotruncana contusa (Cushman) Contusotruncana fornicata (Plummer) Contusotruncana patelliformis (Gandolfi) Contusotruncana plummerae (Gandolfi) Contusotruncana walfischensis (Todd) Costellagerina libyca (Barr) Dicarinella algeriana (Caron) Dicarinella asymetrica (Sigal) Dicarinella canaliculata (Reuss) Dicarinella concavata (Brotzen) Dicarinella hagni (Scheibnerova) Dicarinella imbricata (Mornod) Dicarinella primitiva (Dalbiez) Eohastigerinella watersi (Cushman) Falsotruncana maslakovae Caron Gansserina gansseri (Bolli) Gansserina wiedenmayeri (Gandolfi) Globigerinelloides algerianus Cushman & ten Dam Globigerinelloides alvarezi (Eternod Olvera) Globigerinelloides aptiense Longoria Globigerinelloides barri Bolli, Loeblich & Tappan Globigerinelloides bentonensis (Morrow) Globigerinelloides blowi (Bolli) Globigerinelloides bollii Pessagno Globigerinelloides caseyi (Bolli, Loeblich & Tappan) Globigerinelloides cepedai Longoria Globigerinelloides duboisi (Chevalier) Globigerinelloides ferreolensis Moullade Globigerinelloides gottisi (Chevalier) Globigerinelloides maridalensis (Bolli) Globigerinelloides messinae (Bronnimann) Globigerinelloides prairiehillensis Pessagno Globigerinelloides saundersi (Bolli) Globigerinelloides subcarinatus (Bronnimann) Globigerinelloides ultramicrus (Subbotina) Globotruncana aegyptiaca Nakkady Globotruncana arca (Cushman) Globotruncana bulloides Vogler Globotruncana duwi Nakkady Globotruncana falsostuarti Sigal Globotruncana gagnebini Tilev Globotruncana hilli Pessagno Globotruncana lapparenti Brotzen Globotruncana linneiana (dOrbigny) Globotruncana mariei Banner & Blow Globotruncana orientalis El Naggar Globotruncana rosetta (Carsey) Globotruncana tricarinata (Quereau) Globotruncana ventricosa White Globotruncanella havanensis (Voorwijk) Globotruncanella pschadae (Keller) Globotruncanella petaloidea (Gandolfi) Globotruncanita atlantica (Caron) Globotruncanita conica (White) Globotruncanita elevata (Brotzen) Globotruncanita pettersi (Gandolfi) Globotruncanita stuarti (de Lapparent) Globotruncanita stuartiformis (Dalbiez) Gublerina sp. cf. G. robusta de Klasz Guembelitria sp. Hastigerinoides alexanderi (Cushman) Hedbergella aptiana Bartenstein Hedbergella aptica (Agalarova) Hedbergella delrioensis (Carsey)

323

APPENDIX A. ALPHABETICAL LIST OF ALL CRETACEOUS GENERA AND SPECIES USED FOR PLOTTING SPECIES DIVERSITY
Abathomphalus intermedius (Bolli) Abathomphalus mayaroensis (Bolli) Archaeoglobigerina blowi Pessagno Archaeoglobigerina bosquensis Pessagno Archaeoglobigerina cretacea (dOrbigny) Biglobigerinella multispina Lalicker Biticinella breggiensis (Gandolfi) Biticinella subbreggiensis Sigal Clavihedbergella eocretacea Neagu Clavihedbergella semielongata (Longoria) Clavihedbergella roblesae (Chevalier)

Downloaded from specialpapers.gsapubs.org on March 18, 2011

324
Hedbergella excelsa Longoria Hedbergella flandrini Porthault Hedbergella gorbachikae Longoria Hedbergella hispaniae Longoria Hedbergella holmdelensis Olsson Hedbergella hoelzli (Hagn & Zeil) Hedbergella infracretacea (Glaessner) Hedbergella kuznetsovae (Banner & Desai) Hedbergella luterbacheri Longoria Hedbergella monmouthensis (Olsson) Hedbergella occulta Longoria Hedbergella planispira (Tappan) Hedbergella praetrocoidea Kretzschmar & Gorbachik Hedbergella rhinoceros Coccioni & Cocon Hedbergella rischi Moullade Hedbergella sigali Moullade Hedbergella similis Longoria Hedbergella sliteri Huber Hedbergella trocoidea (Gandolfi) Helvetoglobotruncana helvetica (Bolli) Heterohelix carinata (Cushman) Heterohelix globulosa (Ehrenberg) Heterohelix moremani (Cushman) Heterohelix navarroensis Loeblich Heterohelix punctulata (Cushman) Heterohelix rajagopalani (Govindan) Heterohelix reussi (Cushman) Heterohelix striata (Ehrenberg) Kassabiana falsocalcarata (Kerdany & Abdelsalam) Kuglerina rotundata (Bronnimann) Laeviheterohelix glabrans (Cushman) Laeviheterohelix pulchra (Brotzen) Leupoldina cabri (Sigal) Leupoldina protuberans Bolli Leupoldina pustulans (Bolli) Leupoldina reicheli (Bolli) Marginotruncana coronata (Bolli) Marginotruncana marginata (Reuss) Marginotruncana marianosi (Douglas) Marginotruncana paraconcavata Porthault Marginotruncana pseudolinneiana Pessagno Marginotruncana renzi (Gandolfi) Marginotruncana schneegansi (Sigal) Marginotruncana sigali (Reichel) Marginotruncana sinuosa Porthault Marginotruncana tarfayaensis (Lehmann) Marginotruncana undulata (Lehmann) Planoglobulina acervulinoides (Egger) Planoglobulina carseyae (Plummer) Planomalina buxtorfi (Gandolfi) Planomalina cheniourensis (Sigal) Planomalina praebuxtorfi Wonders Plummerita hantkeninoides (Bronnimann) Praeglobotruncana delrioensis (Plummer) Praeglobotruncana gibba Klaus Praeglobotruncana stephani (Gandolfi) Pseudoguembelina costulata (Cushman) Pseudoguembelina excolata (Cushman) Pseudoguembelina palpebra Bronnimann & Brown Pseudotextularia deformis (Kikoine) Pseudotextularia elegans (Rzehak) Racemiguembelina fructicosa (Egger) Radotruncana calcarata (Cushman) Radotruncana subspinosa (Pessagno) Rotalipora appenninica (Renz)

I. Premoli Silva and W. V. Sliter


Rotalipora balernaensis (Gandolfi) Rotalipora brotzeni (Sigal) Rotalipora cushmani (Morrow) Rotalipora deeckei (Franke) Rotalipora gandolfii Luterbacher & Premoli Silva Rotalipora greenhornensis (Morrow) Rotalipora micheli (Sacal & Debourle) Rotalipora montsalvensis Mornod Rotalipora praeappenninica Sigal Rotalipora praebalernaensis Sigal Rotalipora reicheli Mornod Rotalipora subticinensis (Gandolfi) Rotalipora ticinensis (Gandolfi) Rugoglobigerina hexacamerata Bronnimann Rugoglobigerina macrocephala Bronnimann Rugoglobigerina milamensis Smith & Pessagno Rugoglobigerina pennyi Bronnimann Rugoglobigerina pilula Belford Rugoglobigerina reicheli Bronnimann Rugoglobigerina rugosa (Plummer) Rugotruncana subcircumnodifer (Gandolfi) Schackoina sp. Sigalia carpathica Salaj & Samuel Sigalia decoratissima (de Klasz) Sigalia deflaensis (Sigal) Ticinella bejaouaensis Sigal Ticinella bejaouaensis transitoria Longoria Ticinella madecassiana Sigal Ticinella praeticinensis Sigal Ticinella primula Luterbacher Ticinella raynaudi Sigal Ticinella roberti (Gandolfi) Trinitella scotti Bronnimann Ventilabrella austinana Cushman Ventilabrella eggeri Cushman Ventilabrella glabrata Cushman Ventilabrella multicamerata (de Klasz) Whiteinella aprica (Loeblich & Tappan) Whiteinella archaeocretacea Pessagno Whiteinella aumalensis (Sigal) Whiteinella baltica Douglas & Rankin Whiteinella brittonensis (Loeblich & Tappan) Whiteinella inornata (Bolli) Whiteinella paradubia (Sigal) Whiteinella praehelvetica (Trujillo)

REFERENCES CITED
Arthur, M. A., and Premoli Silva, I., 1982, Development of widespread organic carbon-rich strata in the Mediterranean Tethys, in Schlanger, S. O., and Cita, M. B., eds., Nature and origin of Cretaceous carbon-rich facies: London, Academic Press, p. 754. Arthur, M. A., Schlanger, S. O., and Jenkyns, H. C., 1987, The CenomanianTuronian oceanic anoxic event, II. Paleoceanographic controls on organic matter production and preservation, in Brooks, J., and Fleet, A. J. , eds., Marine petroleum source rocks: Oxford, Geological Society Special Publication, v. 26, p. 371399. Arthur, M. A., Jenkyns, H. C., Brumsack, H. J., and Schlanger, S. O., 1990, Stratigraphy, geochemistry and paleoceanography of organic carbon-rich Cretaceous sequences, in Ginsburg, R. N., and Beaudoin, B., eds., Cretaceous Resources, Events and Rhythms: Dordrecht, Kluwer Academic Publishers, p. 75119. Ascoli, P., 1976, Foraminiferal and ostracod biostratigraphy of the MesozoicCenozoic, Scotian Shelf, Dartmouth, Nova Scotia: Maritime Sediments, Special Publication 1/B, p. 653771.

Downloaded from specialpapers.gsapubs.org on March 18, 2011

Cretaceous paleoceanography: Evidence from planktonic foraminiferal evolution


Bailey, H. W., and Hart, M. B., 1979, The correlation of the early Senonian in Western Europe using Foraminiferida, in Wiedmann, J., ed., Aspekte der Kreide Europas: Stuttgart, International Union of Geological Sciences, Ser. A, v. 6, p. 159169. Banner, F. T., and Desai, D., 1988, A review and revision of the JurassicEarly Cretaceous Globigerinina, with especial reference to the Aptian assemblages of Speeton (north Yorkshire, England): Journal of Micropalaeontology, v. 7, p. 143185. Banner, F. T., Copestake, P., and White, M. R., 1993, Barremian-Aptian Praehedbergellidae of the North Sea area: a reconnaissance: Bulletin of The Natural History Museum, Geology Series, v. 49, p. 130. Barrera, E., and Huber, B. T., 1990, Evolution of Antarctic waters during the Maestrichtian: Foraminifer oxygen and carbon isotope ratios, Leg 113, in Barker, P. F., Kennett, J. P., and others, Proceedings Ocean Drilling Program, Scientific Results, Volume 113: College Station, Texas, Ocean Drilling Program, p. 813827. Baumgartner, P. O., Bown, P., Marcoux, J., Mutterlose, J., Kaminski, M. A., Haig, D. W., and McMinn, A., 1992, Early Cretaceous biogeographic and oceanographic synthesis of Leg 123 (off northwestern Australia), in Gradstein, F. M., Luden, J. N., and others, Proceedings Ocean Drilling Program, Scientific Results, Volume 123: College Station, Texas, Ocean Drilling Program, p. 739758. B, A. W. H., 1977, An Ecological, Zoogeographic and Taxonomic Review of Recent Planktonic Foraminifera, in Ramsay, A. T. S., ed., Oceanic Micropaleontology, Volume 1: London, Academic Press, p. 1100. B, A. W. H., 1980, Gametogenic calcification in a spinose planktonic foraminifer, Globigerinoides sacculifer (Brady): Marine Micropaleontology, v. 5, p. 283310. B, A. W. H., 1982, Biology of planktonic Foraminifera, in Broadhead, T. W., ed., Foraminifera: Notes for a Short Course, Studies in Geology, v. 6, p. 5192. Bellanca, A., Claps, M., Erba, E., Masetti, D., Neri, R., Premoli Silva, I., and Venezia, F., 1996, Orbitally induced limestone-marlstone rhythms in the Albian-Cenomanian Cismon section (Venetian region, northern Italy): sedimentology, calcareous and siliceous plankton distribution, elemental and isotope geochemistry: Palaeogeography, Palaeoclimatology, Palaeoecolology, v. 126, p. 227260. Berggren, W. A., 1962, Some planktonic Foraminifera from the Maestrichtian and type Danian Stages of Southern Scandinavia: Stockholm Contributions in Geology, v. 9, no. 1, 106 p. Boersma, A., and Premoli Silva, I., 1983, Paleocene planktonic foraminiferal biogeography and paleoceanography of the Atlantic Ocean: Micropaleontology, v. 29, p. 355381. Boersma, A., and Premoli Silva, I., 1988, Atlantic Paleogene biserial heterohelicid Foraminifera and oxygen minima: Paleoceanography, v. 4, p. 271286. Boersma, A., and Premoli Silva, I., 1991, Distribution of Paleogene planktonic Foraminifera analogies with the Recent?: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 83, p. 2948. Boersma, A., and Shackleton, N. J., 1981, Oxygen and carbon isotope variations and planktonic foraminiferal depth habitats: Late Cretaceous to Paleocene, central Pacific, DSDP Sites 463 and 465, Leg 62, in Thiede, J., Vallier, T. L., and others, Initial Reports of the Deep Sea Drilling Project, Volume 62: Washington, D.C., U.S. Government Printing Office, p. 513526. Boersma, A., Premoli Silva, I., and Shackleton, N. J., 1987, Atlantic Eocene planktonic foraminiferal paleohydrographic indicators and stable isotope paleoceanography: Paleoceanography, v. 2, p. 287331. Brhret, J. G., Caron, M., and Delamette, M., 1986, Niveaux riches en matire organique dans lAlbien vocontien; quelques caractres du paloenvironnement; essai dinterpretation gntique (Layers rich in organic matter from the Vocontian Albian; some paleoenvironmental characteristics; a tentative of genetic interpretation): Documents du Bureau de Recherches Gologiques et Minires, no. 110, p. 141191. Caron, M., 1981, Un nouveau genre de foraminifre planctonique du Crtac: Falsotruncana nov. gen. (The new Cretaceous planktonic foraminiferal genus: Falsotruncana n. gen.): Eclogae geologicae Helvetiae, v. 74, p. 6573.

325

Caron, M., 1983, La spciation chez les Foraminifres planctiques: une rponse adapte aux contraintes de lenvironnement (Speciation within planktonic Foraminifera: a response to environmental constraints): Zitteliana, v. 10, p. 671676. Caron, M., 1985, Cretaceous planktic Foraminifera, in Bolli, H. M., Saunders, J. B., and Perch-Nielsen, K., eds., Plankton Stratigraphy: Cambridge, Cambridge Earth Sciences Series, Cambridge University Press, p. 1786. Caron, M., and Homewood, P., 1983, Evolution of early planktic foraminifers: Marine Micropaleontology, v. 7 (1982/83), p. 453462. Cecca, F., Pallini, G., Erba, E., Premoli Silva, I., and Coccioni, R., 1994, Hauterivian-Barremian chronostratigraphy based on ammonites, nannofossils, planktonic Foraminifera, and magnetic chrons from Mediterranean Domain: Cretaceous Research, v. 15, p. 457467. Channell, J. E. T., Erba, E., Nakanishi, M., and Tamaki, K., 1995, Late JurassicEarly Cretaceous Time Scales and oceanic magnetic anomaly block models, in Berggren, W. A., Kent, D. V., Aubry, M. P., and Hardenbol, J., eds., Geochronology time scales and global stratigraphic correlation: Society of Economic Paleontologists and Mineralogists Special Publication, no. 54, p. 5163. Ciesielski, P. F., Kristoffersen, Y., and others, 1988, Proceedings Ocean Drilling Program, Initial Reports, Leg 114: College Station , Texas, Ocean Drilling Program, 815 p. Coccioni, R., and Premoli Silva, I., 1994, Planktonic Foraminifera from the Lower Cretaceous of Rio Argos sections (southern Spain) and biostratigraphic implications: Cretaceous Research, v. 15, p. 645687. Coccioni, R., Erba, E., and Premoli Silva, I., 1991, Litho- and biostratigraphy of the Livello Bonarelli close to the Cenomanian/Turonian boundary (Umbria-Marche Apennines, Italy) and possible paleoceanographic significance: Colloque International sur les Evnements de la limite Cnomanien-Turonien, Grenoble 1991, Gologie Alpine, Mmoire Hors Srie, v. 17, p. 2526. Coccioni, R., Erba, E., and Premoli Silva, I., 1992, Barremian-Aptian calcareous plankton biostratigraphy from the Gorgo Cerbara section (Marche, central Italy) and implications for plankton evolution: Cretaceous Research, v. 13, p. 517537. Davids, R. N., 1966, A paleoecologic and paleo-biogeographic study of Maastrichtian planktonic Foraminifera [Ph.D. thesis]: New Brunswick, New Jersey, Rutgers State University, 241 p. Delacotte, O., Renard, M., Laj, C., Perch-Nielsen, K., Premoli Silva, I., and Clauser, S., 1985, Magntostratigraphie et biostratigraphie du passage Crtac-Tertiaire de la coupe de Bidart (Pyrnes Atlantiques) (Magnetostratigraphy and biostratigraphy of the Cretaceous-Tertiary transition from the Bidart section [Atlantic Pyrenees]): Gologie de la France, no. 3, p. 243253. De Mir, M. D., 1971, Los foraminiferos planctnicos vivos y sedimentados del margen continental de Venezuela (resumen) (Living and sedimented planktonic foraminifers from the continental margin of Venezuela [Abstract]): Acta Geologica Hispanica, v. 6, p. 102108. Dercourt, J., Ricou, L. E., and Vrielynck, B., editors, 1993, Atlas Tthys Paleoenvironmental Maps. Explanatory Notes: Paris, Gauthier-Villars, 307 p. DHondt, S., and Arthur, M. A., 1995, Interspecific variation in stable isotope signals of Maastrichtian planktonic Foraminifera: Paleoceanography, v. 10, p. 123135. DHondt, S., and Arthur, M. A., 1996, Late Cretaceous and the cool tropic paradox: Science, v. 271, p. 18381841. Douglas, R. G., 1969, Upper Cretaceous planktonic Foraminifera in northern California. Part 1-Systematics: Micropaleontology, v. 15, p. 151209. Douglas, R. G., and Rankin, C., 1969, Cretaceous planktonic Foraminifera from Bornholm and their zoogeographic significance: Lethaia, v. 2, p. 185217. Douglas, R. G., and Savin, S. M., 1975, Oxygen and carbon isotope analyses of Tertiary and Cretaceous microfossils from the Shatsky Rise and other sites in the North Pacific Ocean, in Larson, R. L., Moberly, R., and others, Initial Reports of the Deep Sea Drilling Project, Volume 32: Washington, D.C., U.S. Government Printing Office, p. 509520. Elder, W. P., 1989, Molluscan extinction patterns across the Cenomanian-Turo-

Downloaded from specialpapers.gsapubs.org on March 18, 2011

326

I. Premoli Silva and W. V. Sliter


Foraminiferida from the Cenomanian of Bornholm, Denmark: Newsletters in Stratigraphy, v. 8, p. 8396. Hart, M. B., 1980, A water depth model for the evolution of the planktonic Foraminiferida: Nature, v. 286, p. 252254. Hart, M. B., and Bailey, H. W., 1979, The distribution of planktonic Foraminiferida in the mid-Cretaceous of NW Europe, in Wiedmann, J., ed., Aspekte der Kreide Europas: Stuttgart, International Union of Geological Sciences, Ser. A, v. 6, p. 527-542. Hart, M. B., and Ball, K. C., 1986, Late Cretaceous anoxic events, sea-level changes and the evolution of the planktonic Foraminifera, in Summerhayes, C. P., and Shackleton, N. J., eds., North Atlantic Paleoceanography: Geological Society of London, Special Publication no. 21, p. 6778. Hart, M. B., and Tarling, D. H., 1974, Cenomanian palaeogeography of the North Atlantic and possible mid-Cenomanian eustatic movements and their implications: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 15, p. 95108. Hemleben, C., Spindler, M., and Anderson, O. R., 1989, Modern planktonic Foraminifera: Berlin, Springer-Verlag, 363 p. Herb, R., 1974, Cretaceous planktonic Foraminefera from the eastern Indian Ocean, in Davies, T. A., Luyendyk, B. P., and others, Initial reports of the Deep Sea Drilling Project, Volume 26: Washington, D.C., U.S. Government Printing Office, p. 745769. Herbert, T. D., and Fischer, A. G., 1986, Milankovitch climatic origin of mid-Cretaceous black shale rhythms in central Italy: Nature, v. 321, p. 739743. Herbert, T. D, Premoli Silva, I., Erba, E., and Fischer, A. G., 1995, Orbital chronology of Cretaceous-Paleocene marine sediments, in Berggren, W. A., Kent, D. V., Aubry, M.P., and Hardenbol, J., eds., Geochronology Time Scales and Global Stratigraphic Correlation: Society of Economic Paleontologists and Mineralogists, Special Publication no. 54, p. 8193. Herbin, J. P., Montardert, L., Mueller, C., Gomez, R., Thurow, J., and Wiedman, J., 1986, Organic-rich sedimentation at the Cenomanian-Turonian boundary in oceanic and coastal basins in the North Atlantic and Tethys, in Summerhayes, C. P., and Shackleton, N. J., eds., North Atlantic paleoceanography: Geological Society of London, Special Publication, no. 21, p. 389422. Hilbrecht, H., Hubberten, H.W., and Oberhaensli, H., 1992, Biogeography of planktonic Foraminifera and regional carbon isotope variations: productivity and water masses in Late Cretaceous Europe: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 92, p. 407421. Huber, B. T., 1988, Upper CampanianPaleocene Foraminifera from the James Ross Island region, Antarctic Peninsula: Geological Society of America, Memoire 169, p. 163252. Huber, B. T., 1990, Maastrichtian planktonic foraminifer biostratigraphy of the Maud Rise, ODP Leg 113 Sites 689 and 690, Weddell Sea, in Barker, P. F., Kennett, J. P., and others, Proceedings Ocean Drilling Program, Scientific Results, Volume 113: College Station, Texas, Ocean Drilling Program, p. 319324. Huber, B. T., 1991, Planktonic foraminifer biostratigraphy of CampanianMaestrichtian sediments from Sites 698 and 700, southern South Atlantic, in Ciesielski, P. F., Kristoffersen, Y., and others, Proceedings Ocean Drilling Program, Scientific Results, Volume 114: College Station, Texas, Ocean Drilling Program, p. 281297. Huber, B. T., 1992, Paleobiogeography of Campanian-Maastrichtian foraminifers in the southern high latitudes: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 92, p. 325360. Huber, B. T., and Boersma, A., 1994, Cretaceous origination of Zeauvigerina and its relationship to Paleocene biserial planktonic Foraminifera: Journal of Foraminiferal Research, v. 24, p. 268287. Huber, B. T., and Watkins, D. K., 1992, Biogeography of Campanian-Maastrichtian calcareous plankton in the region of the southern ocean: paleogeographic and paleoclimatic implications: The Antarctic paleoenvironment: A Perspective on Global Change, Antarctic Research Series, v. 56, p. 3145. Huber, B. T., Hodell, D. A., and Hamilton, C. P., 1995, MiddleLate Cretaceous climate of the southern high latitudes: Stable isotopic evidence for minimal equator-to-pole thermal gradients: Geological Society of America

nian Stage boundary in the western interior of the United States: Paleobiology, v. 15, p. 299320. Erba, E., 1992, Calcareous nannofossil distribution in pelagic rhythmic sediments (Aptian-Albian Piobbico core, central Italy): Rivista Italiana di Paleontologia e Stratigrafia, v. 97 (1991), p. 455484. Erba, E., 1994, Nannofossils and superplumes: The early Aptian nannoconid crisis: Paleoceanography, v. 9, p. 483501. Erba, E., and Quadrio, B., 1987, Biostratigrafia a Nannofossili calcarei, Calpionellidi e Foraminiferi planctonici della Maiolica (Titoniano superiore-Aptiano) nelle Prealpi Bresciane (Italia settentrionale) (Biostratigraphy of calcareous nannofossils, calpionellids, and planktonic foraminifers of the Maiolica Formation [upper Tithonian-Aptian] from Brescian Prealps [northern Italy]): Rivista Italiana di Paleontologia e Stratigrafia, v. 93, p. 3108. Erba, E., Premoli Silva, I., and Watkins, D. K., 1995a, Cretaceous calcareous plankton biostratigraphy of Sites 871 through 879 (Leg 144), in Haggerty, J. A., Premoli Silva, I., Rack, F., and McNutt, M. K., eds., Proceedings Ocean Drilling Program, Scientific Results, Volume 144: College Station, Texas, Ocean Drilling Program, p. 157169. Erba, E., Premoli Silva, I., Wilson, P. A., Pringle, M. S., Sliter, W. V., Watkins, D. K., Arnaud Vanneau, A., Bralower, T. I., Budd, A. F., Camoin, G. F., Masse, J.P., Mutterlose, J., and Sager, W. V., 1995b, Synthesis of stratigraphies from shallow-water sequences at Sites 871 through 879 in the Western Pacific Ocean (Leg 144), in Haggerty, J. A., Premoli Silva, I., Rack, F., and McNutt, M. K., eds., Proceedings Ocean Drilling Program, Scientific Results, Volume 144: College Station, Texas, Ocean Drilling Program, p. 873885. Erbacher, J., 1994, Entwicklung und Palaeoozeanographie mittelkretazischer Radiolarien der westlichen Tethys (Italien) und des Nordatlantiks (Development and paleoceanography of mid-Cretaceous radiolarians): Tuebinger Mikropalaeontologische Mitteilungen, v. 12, 120 p. Fairbanks, R. G., and Wiebe, P. H., 1980, Foraminifera and chlorophyll maximum: Vertical distribution, seasonal succession, and paleoceanographic significance: Science, v. 207, p. 6163. Golonka, J., Ross, M. I., and Scotese, C. R., 1994, Phanerozoic paleoceanographic and paleoclimatic modeling maps, in Embry, A. F., Beauchamp, B., and Glass, J., eds., Pangea: Global environments and resources: Canadian Society of Petroleum Geologists, Memoire 17, p. 147. Gradstein, F. M., Ludden, J. N., and others, 1990, Proceedings Ocean Drilling Program, Initial Reports, Volume 123: College Station, Texas, Ocean Drilling Program, 716 p. Gradstein, F. M., Agterberg, F. P., Ogg, J. G., Hardenbol, J., van Veen, P., Thierry, J., and Huang, Z., 1994, A Mesozoic time scale: Journal of Geophysic Research, v. 99, no. B12, p. 24,05124,074. Haggerty, J. A., Premoli Silva, I., Rack, F. R., and McNutt, M. K, editors, 1995, Proceedings of Ocean Drilling Program, Scientific Results, Volume 144: College Station, Texas, Ocean Drilling Program, 1059 p. Haig, D. W., 1992, Aptian-Albian foraminifers from Site 766, Cuvier Abyssal Plain, and comparison with coeval faunas from the Australian region, in Gradstein, F. M., Ludden, J. N., and others, Proceedings Ocean Drilling Program, Scientific Results, Volume 123: College Station, Texas, Ocean Drilling Program, p. 271297. Hallock, P., 1985, Why are larger Foraminifera large?: Paleobiology, v. 11, p. 195208. Hallock, P., 1987, Fluctuations in the trophic resource continuum: A factor in global diversity cycles?: Paleoceanography, v. 2, p. 457471. Hallock, P., and Schlager, W., 1986, Nutrient excess and the demise of coral reefs and carbonate platforms: Palaios, v. 1, p. 389398. Hallock, P., Premoli Silva, I., and Boersma, A., 1991, Similarities between planktonic and larger foraminiferal evolutionary trends through Paleogene paleoceanographic changes: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 83, p. 4964. Hart, M. B., 1976, The mid-Cretaceous succession of Orphan Knoll (northwest Atlantic): micropaleontology and palaeo-oceanographic implications: Canadian Journal of Earth Sciences, v. 13, p. 14111421. Hart, M. B., 1979, Biostratigraphy and paleozoogeography of planktonic

Downloaded from specialpapers.gsapubs.org on March 18, 2011

Cretaceous paleoceanography: Evidence from planktonic foraminiferal evolution


Bulletin, v. 107, p. 11641191. Keller, G., 1993, The Cretaceous-Tertiary boundary transition in the Antarctic Ocean and its global implications: Marine Micropaleontology, v. 21, p. 145. Keller, G., Barrera, E., Schmitz, B., and Mattson, E., 1993, Gradual mass extinction, species survivorship, and long-term environmental changes across the Cretaceous-Tertiary boundary in high latitudes: Geological Society of America Bulletin, v. 105, p. 979997. Kennett, J. P., 1982, Marine geology: Englewood Cliffs, New Jersey, PrenticeHall, 813 p. Krasheninnikov, V. A., and Basov, I. A., 1983, Stratigraphy of Cretaceous sediments of the Falkland Plateau based on planktonic foraminifers, Deep Sea Drilling Project, Leg 71, in Ludwig, W. J., Krasheninnikov, V. A., and others, Initial reports of the Deep Sea Drilling Project, Volume 71: Washington, D.C., U.S. Government Printing Office, p. 789820. Kroon, D., and Ganssen, G., 1988, Northern Indian Ocean upwelling cells and the stable isotope composition of living planktic Foraminifera, in Brummer, G., and Kroon, D., eds., Planktonic Foraminifera as Tracers of Ocean-Climate History: Amsterdam, Free University Press, p. 299319. Kroon, D., and Nederbragt, A. J., 1990, Ecology and paleoecology of triserial planktic foraminifera: Marine Micropaleontology, v. 16, p. 2538. Larson, R. L., 1991, Geological consequences of superplumes: Geology, v. 19, p. 963966. Laughton, A. S., Berggren, W. A., and others, 1972, Initial reports of the Deep Sea Drilling Project, Volume 12: Washington, D.C., U.S. Government Printing Office, 1243 p. Leckie, R. M., 1987, Paleoecology of mid-Cretaceous planktonic Foraminifera: A comparison of open ocean and epicontinental sea assemblages: Micropaleontology, v. 33, p. 164176. Leckie, R. M., 1989, A paleoceanographic model for the early evolutionary history of planktonic Foraminifera: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 73, p. 107138. Leckie, R. M., 1990, Middle Cretaceous planktonic foraminifers of the Antarctic margin: Hole 693A, ODP Leg 113, in Barker, P. F., Kennett, J. P., and others, Proceedings Ocean Drilling Program, Scientific Results, Volume 113: College Station, Texas, Ocean Drilling Program, p. 319324. Leereveld, H., 1995, Dinoflagellate cysts from the Lower Cretaceous Rio Argos succession (SE Spain): Utrecht, Laboratory of Palaeobotany and Palynology Foundation, Laboratory of Palaeobotany and Palynology, Contribution Series, no. 2, 171 p. Lini, A., Weissert, H., and Erba, E., 1992, The Valanginian carbon isotope event: a first episode of greenhouse climate conditions during the Cretaceous, in Wezel, F. C., ed., Global change special issue: Terra Nova, v. 4, p. 374384. Longoria, J. F., 1974, Stratigraphic, morphologic and taxonomic studies of Aptian planktonic Foraminifera: Revista Espanola de Micropaleontologia, Numero Extraordinario, 150 p. MacArthur, R. H., and Wilson, E. O., 1967, The theory of island biogeography: Princeton, New Jersey, Princeton University Press, 167 p. Magniez-Jannin, F., 1981, Decouverte de Planomalina buxtorfi (Gandolfi) et d autres Foraminifres planctoniques inattendus dans lAlbien suprieur d Abbotscliff (Kent, Angleterre); Consequences palogographiques et biostratigraphiques (Unexpected occurrence of Planomalina buxtorfi and other planktonic Foraminifera from the upper Albian of Abbotscliff [Kent, England]; paleogeographic and biostratigraphic implications): Gobios, v. 14, p. 9197. Marcucci Passerini, M., Bettini, P., Dainelli, J., and Sirugo, A., 1991, The Bonarelli Horizon in the central Apennines (Italy): radiolarian biostratigraphy: Cretaceous Research, v. 12, p. 321331. Margalef, R., 1965, Composicin y distributin del fitoplancton (Composition and distribution of phytoplankton): Memoria de la Sociedad Ciencias Naturales La Salle, v. 25, p. 141205. Mutterlose, J., 1989, Temperature controlled migration of calcareous nannofloras in the NW European Aptian, in Crux, J. A., and van Heck, S. E., eds., Nannofossils and their applications: Chichester, Ellis Horwood, p. 122142.

327

Mutterlose, J., 1991, Das Verterilungs- und Migrations-Muster des kalkigen Nannoplanktons in der Unterkreide (Valangin-Apt) NW-Deutschland (Distribution and migration patterns of calcareous nannoplankton in the Lower Cretaceous [Valanginian-Aptian] from NW Germany): Palaeontographica, Abt. B, v. 221, p. 27152. Nederbragt, A. J., 1991, Late Cretaceous biostratigraphy and development of Heterohelicidae (planktic foraminifera): Micropaleontology, v. 37, p. 329372. Olsson, R. K., Hemleben, C., Berggren, W. A., and Liu, C., 1992, Wall texture classification of planktonic Foraminifera Genera in the lower Danian: Journal of Foraminiferal Research, v. 22, p. 195213. Pessagno, E. A., 1967, Upper Cretaceous planktonic Foraminifera from the western Gulf Coastal Plain: Paleontographica Americana, v. 5, p. 245445. Premoli Silva, I., 1992, Oldest Cretaceous planktonic foraminifers from Hole 700B, in Ciesielski, P. F., Kristoffersen, Y., and others, Proceedings Ocean Drilling Program, Scientific Results, Volume 114: College Station, Texas, Ocean Drilling Program, p. 281297. Premoli Silva, I., and Boersma, A., 1989, Atlantic Paleogene planktonic foraminiferal bioprovincial Indices: Marine Micropaleontology, v. 14, p. 357371. Premoli Silva, I., and Sliter, W. V., 1994, Cretaceous planktonic foraminiferal biostratigraphy and evolutionary trends from the Bottaccione section, Gubbio, Italy: Palaeontographia Italica, v. 82 (1995), p. 189. Premoli Silva, I., Erba, E., and Tornaghi, M. E., 1989a, Paleoenvironmental signals and changes in surface fertility in Mid Cretaceous Corg-rich pelagic facies of the Fucoid Marls (central Italy): Gobios, Memoire Special, no. 11, p. 225236. Premoli Silva, I., Ripepe, M., and Tornaghi, M. E., 1989b, Planktonic foraminiferal distribution records productivity cycles: evidence from the AptianAlbian Piobbico core (central Italy): Terra Nova, v. 1, p. 443448. Premoli Silva, I., Garzanti, E., and Gaetani, M., 1991, Stratigraphy of the Chikkim and Fatu La Formations in the Zangla and Zumlung Units (Zanskar Range, India) with comparison to the Thakkhola region (central Nepal): Mid-Cretaceous evolution of the Indian passive margin: Rivista Italiana di Paleontologia e Stratigrafia, v. 97, p. 511564. Rhyther, J. H., 1969, Photosynthesis and fish production in the sea: Science, v. 166, p. 7276. Robaszynski, F., and Caron, M., 1995, Foraminifres planctoniques du Crtac: commentaire de la zonation Europe-Mditerrane (Cretaceous planktonic Foraminifera: comments on the European-Mediterranean zonation): Bulletin de la Socit Gologique de France, v. 166, no. 6, p. 681692. Robaszynski, F., Caron, M., editors, and The European Working Group on Planktonic Foraminifera, 1979, Atlas of Mid Cretaceous planktonic Foraminifera (Boreal Sea and Tethys). Parts 12: Cahiers de Micropalontologie, 366 p., 80 pl. Robaszynski, F., Caron, M., Gonzales Donoso, J. M., Wonders, A. H., editors, and The European Working Group on Planktonic Foraminifera, 1984, Atlas of Late Cretaceous Globotruncanids: Revue de Micropalontologie, v. 26, p. 145305, 54 pl. Robaszynski, F., Caron, M., Dupuis, C., Amedro, F., Gonzales-Donoso, J. M., Linares, D., Harbenbol, J., Gartner, S., Calendra, F., and Deloffre, R., 1990, A tentative integrated stratigraphy in the Turonian of central Tunisia: formations, zones and sequential stratigraphy in the Kalaat Senan area: Bulletin des Centres de Recherches Exploration-Production ElfAquitaine, v. 14, p. 213384. Robaszynski, F., Hardenbol, J., Caron, M., Amedro, F., Dupuis, C., GonzalesDonoso, J. M, Linares, D., and Gartner, S., 1993, Sequence stratigraphy in a distal environment: the Cenomanian of the Kalaat Senan region (central Tunisia): Bulletin des Centres de Recherches Exploration-Production Elf-Aquitaine, v. 17, p. 395433. Roth, P. H., 1978, Cretaceous nannoplankton biostratigraphy and oceanography of the northwestern Atlantic Ocean, in Benson, W. E., Sheridan, R. E., and others, Initial reports of the Deep Sea Drilling Project, Volume 44: Washington, D.C., U.S. Government Printing Office, p. 731759. Roth, P. H., and Krumbach, K. R., 1986, Middle Cretaceous calcareous nannofossil biogeography and preservation in the Atlantic and Indian Oceans:

Downloaded from specialpapers.gsapubs.org on March 18, 2011

328

I. Premoli Silva and W. V. Sliter


raphy: Abhadlungen der Geologischen Bundesanstalt, v. A 29, p. 152. Thurow, J., Brumsack, H. J., Littke, R., Meyers, P., and Rullkoetter, J., 1992, Mid-Cretaceous events in the Indian Ocean a key to understanding the global picture, in Duncan, R. A., Rea, D. K., Kidd, R. B., von Rad, U., and Weissel, J. K., eds., Synthesis of results from Scientific Drilling in the Indian Ocean: Washington, D.C., American Geophysical Union, Geophysical Monograph, v. 70, p. 253273. Tornaghi, M. E., Premoli Silva, I., and Ripepe, M., 1989, Lithostratigraphy and planktonic foraminiferal biostratigraphy of the Aptian-Albian Scisti a Fucoidi in the Piobbico core, Marche, Italy: background for cyclostratigraphy: Rivista Italiana di Paleontologia e Stratigrafia, v. 95, p. 223264. Valentine, J. W., 1973, Evolutionary ecology of the marine biosphere: Englewood Cliffs, New Jersey, Prentice-Hall, 511 p. Venkatachalapathy, R., and Ragothaman, V., 1995, A foraminiferal zonal scheme for the mid-Cretaceous sediments of the Cauvery Basin, India: Cretaceous Research, v. 16, p. 415433. Weissert, H., 1989, C-isotope stratigraphy, a monitor of paleoenvironmental change: a case study from the Early Cretaceous: Surveys in Geophysics, v. 10, p. 1-61. Weissert, H., and Brhret, J. G., 1991, A carbonate carbon-isotope record from the Aptian-Albian sediments of the Vocontian trough (SE France): Bulletin de la Socit Gologique de France, v. 162, p. 11331140. Weissert, H., and Lini, A., 1991, Ice age interlude during the time of Cretaceous greenhouse climate?, in Mller, D. W., McKenzie, J. A., and Weissert, H., eds., Controversies in Modern Geology: London, Academic Press, p. 173191. Winterer, E. L., Sager, W. W., Firth, J. V., and Sinton, J. M., eds., 1995, Proceedings of Ocean Drilling Program, Scientific Results, Volume 143: College Station, Texas, Ocean Drilling Program, 629 p. Wonders, A. A. H., 1979, Middle and Late Cretaceous pelagic sediments of the Umbrian Sequence in the Central Apennines: Proceedings, Koninklijke Nederlandse Akademie van Wetenschappen, B, v. 82, p. 171205. Wonders, A. A. H., 1980, Middle and Late Cretaceous planktonic Foraminifera on the western Mediterranean area: Utrecht Micropaleontological Bulletin, no. 24, 157 p. Wonders, A. A. H., 1992, Cretaceous planktonic foraminiferal biostratigraphy, Leg 122, Exmouth Plateau, Australia, in von Rad, U., Haq, B. U., and others, Proceedings Ocean Drilling Program, Scientific Results, Volume 122: College Station, Texas, Ocean Drilling Program, p. 587599. Zimmerman, H. B., Boersma, A., and McKoy, F. W., 1987, Carbonaceous sediments and palaeoenvironment of the Cretaceous South Atlantic Ocean, in Brooks, J., and Fleet, A. J., eds., Marine Petroleum Source Rocks: Oxford, Geological Society Special Publication, v. 26, p. 271286. MANUSCRIPT ACCEPTED BY THE SOCIETY JULY 14, 1998

implications for paleoceanography: Marine Micropaleontology, v. 10, p. 235266. Schlanger, S. O., 1986, High frequency sea-level fluctuations in Cretaceous time: an emerging geophysical problem, in Hsu, K. J., ed., Mesozoic and Cenozoic oceans: American Geophysical Union Geodynamic Series, v. 15, p. 6174. Sigal, J., 1977, Essai de zonation du Crtac mditerranen l aide des foraminifres planctoniques (Planktonic foraminiferal zonation of the Mediterranean Cretaceous): Gologie Mditerranenne, v. 4, p. 99108. Sissingh, W., 1977, Biostratigraphy of Cretaceous calcareous nannoplankton: Geologie en Mijnbouw, v. 56, p. 3765. Sliter, W. V., 1972, Upper Cretaceous planktonic foraminiferal zoogeography and ecologyEastern Pacific Margin: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 12, p. 1531. Sliter, W. V., 1976, Cretaceous foraminifers from the southwestern Atlantic Ocean, Leg 36, Deep Sea Drilling Project, in Barker, P. F., Dalziel, I. D. W., and others, Initial reports of the Deep Sea Drilling Project, Volume 36: Washington, D.C., U.S. Government Printing Office, p. 519545. Sliter, W. V., 1980, Mesozoic foraminifers and deep-sea benthic environments from Deep Sea Drilling Project Sites 415 and 416, eastern North Atlantic, in Lancelot, Y., Winterer, E. L., and others, Initial reports of the Deep Sea Drilling Project, Volume 50: Washington, D.C., U.S. Government Printing Office, p. 353427. Sliter, W. V., 1989, Aptian anoxia in the Pacific Basin: Geology, v. 17, p. 909912. Sliter, W. V., 1992, Cretaceous planktonic foraminiferal biostratigraphy and paleoceanographic events in the Pacific Ocean with emphasis on indurated sediments, in Ishizaki, K., and Saito, T., eds., Centenary of Japanese micropaleontology: Tokyo, Terra Scientific Publishing Company, p. 281299. Sliter, W. V., 1995, Cretaceous planktonic foraminifers from Sites 865, 866, and 869: A synthesis of Cretaceous pelagic sedimentation in the central Pacific Ocean Basin, in Winterer, E. L., Sager, W. W., Firth, J. V., and Sinton, J. M., eds., Proceedings Ocean Drilling Program, Scientific Results, Volume 143: College Station, Texas, Ocean Drilling Program, p. 1530. Stenestad, E., 1969, The genus Heterohelix Ehrenberg, 1983 (Foraminifera) from the Senonian of Denmark: Proceedings, First International Conference on Planktonic Microfossils, Geneva 1967, v. 2, p. 644663. Streel, M., Normand, S., Felder, S., and Keppens, E., 1995, Pollen response to short term climatic changes in the late Maastrichtian at ENCI, South Limburg, The Netherlands: Second International Symposium on Cretaceous Stage Boundaries, Brussels, 814 Sept. 1995, Abstract Volume, p. 117. Tappan, H., 1962, Foraminifera from the Arctic Slope of Alaska, Part 3, Cretaceous Foraminifera: Geological Survey Professional Paper 236C, p. 91209. Thierstein, H. R., 1973, Lower Cretaceous calcareous nannoplankton biostratig-

Printed in U.S.A.

Вам также может понравиться