<<

Earth and Planetary Science Letters 234 (2005) 299–315 www.elsevier.com/locate/epsl Frontiers The role of the global cycle in the regulation and evolution of the Earth system

Andy Ridgwella,T, Richard E. Zeebeb

aDepartment of Earth and Sciences, The University of British Columbia, 6339 Stores Road, Vancouver, British Columbia, Canada V6T 1Z4 bUniversity of Hawaii at Manoa, SOEST, Honolulu, HI, United States Received 8 October 2004; received in revised form 7 March 2005; accepted 17 March 2005 Available online 27 April 2005

Abstract

We review one of the most ancient of all the global biogeochemical cycles and one which reflects the profound geochemical and biological changes that have occurred as the Earth system has evolved through time—that of carbonate (CaCO3). In particular, we highlight a Mid- Revolution in the and location of carbonate in the marine environment, driven by the ecological success of . This drove the creation of a responsive deep- sedimentary sink of CaCO3. The result is that biologically driven carbonate deposition provides a significant buffering of ocean and of atmospheric CO2 in the modern system. However, the same calcifying that under-pin the deep-sea carbonate sink are now threatened by the continued atmospheric release of fuel CO2 and increasing acidity of the surface ocean. We are not yet in a position to predict what the impact on CaCO3 production will be, or how the uptake of CO2 by the ocean will be affected. This uncertainty in the future trajectory of atmospheric CO2 that comes from incomplete understanding of the marine carbonate cycle is cause for concern. D 2005 Elsevier B.V. All rights reserved.

Keywords: Earth system; cycle; carbonate; calcifiers; ; CO2; fossil fuel

1. Introduction , and exchangeable in the marine environment (Fig. 1) while the latter include crustal The geochemical or long-term rocks and deeply buried sediments in addition to the primarily involves the exchange of carbon between underlying mantle. How carbon is partitioned the dsurficialT and dgeologicT reservoirs [1]. The between the various reservoirs of the surficial system former comprise , , , and between surficial and geologic reservoirs is what sets the concentration of CO2 in the atmosphere. T Corresponding author. Tel.: +1 604 822 2449; fax: +1 604 822 Life, and the cycle of organic carbon, as well as its 6088. geological (and subsequent fossil fuel exhumation) is E-mail address: [email protected] (A. Ridgwell). of particular importance in this regard. The cycle of

0012-821X/$ - see front matter D 2005 Elsevier B.V. All rights reserved. doi:10.1016/j.epsl.2005.03.006 300 A. Ridgwell, R.E. Zeebe / Earth and Planetary Science Letters 234 (2005) 299–315

Box 1 Carbonate chemistry d101T and jargon buster

2+ The (CaCO3) has a crystal lattice motif comprising one calcium (Ca ) 2À ionically bound to one carbonate ion (CO3 ), configured in different polymorphic forms; e.g., ,a trigonal structure, or , which is orthorhombic. Precipitation may be described by the following 2+ À 2+ reaction: Ca +2HCO3 YCaCO3 +CO2(aq) +H2O. Of the reactants required for this, Ca is naturally abundant in sea- and at one of the highest concentrations of all ionic species in the ocean. À (HCO3 ) are also ubiquitous in sea-water and are formed through the dissolution of CO2 gas. Under typical marine conditions, will largely hydrate to form a proton (H+) and a + À bicarbonate ion (HCO3); H2O+CO2(aq)YH +HCO3 (see Fig. 3), while true carbonic (H2CO3)is À only present in very small concentrations. A fraction of HCO3 dissociates to form a carbonate ion 2À À + 2À À 2À (CO3 ); HCO3 YH +CO3 . The sum total; CO2(aq) (+H2CO3)+HCO3 +CO3 is collectively termed dissolved inorganic carbon (dDICT). The climatic importance of the CaCO3 precipitation reaction arises because although the sum total of dissolved carbon species (DIC) is reduced, the remaining carbon is re-partitioned in favor of CO2(aq), resulting in a higher of CO2 ( pCO2) in the surface ocean. (Another way of thinking 2À about this is in terms of removing CO3 and shifting the aqueous carbonate equilibrium reaction 2À À CO2(aq) +CO3 +H2O X 2HCO3 to the left to compensate.) The counter-intuitive and often confusing consequence of all this is that the precipitation of carbonate carbon drives an increase in ocean pCO2, and with it, an increase in atmospheric CO2 concentration. Conversely, dissolution of CaCO3 drives a pCO2 (and atmospheric CO2) decrease. Whether CaCO3 precipitates or dissolves depends on the relative stability of its . This 2+ 2À can be directly related to the ambient concentrations (strictly, activities) of Ca and CO3 by the 2+ 2À saturation state (also known as the ratio) X of the solution, defined; X =[Ca ]Â[CO3 ]/Ksp, where Ksp is a solubility constant [44]. The precipitation of calcium carbonate from sea-water is thermodynamically favorable when X is greater than unity and occurs at a rate taking the form of a proportionality with (XÀ1)n [100], where n is a measure of how strongly the precipitation rate responds to 2À n a change in CO3 . Conversely, CaCO3 will tend to dissolve at X b1.0, and at a rate proportional to (1ÀX) [101]. 2+ 2À As well as the concentrations of Ca and CO3 , depth in the ocean is also important because Ksp scales with increasing pressure. Since Ksp and X are inversely related, the greater the depth in the ocean the more likely the ambient environment is to be under-saturated (i.e., X b1.0). The depth at which X =1.0 occurs is termed the equilibrium calcite saturation horizon (CSH). (Similar terminology can be applied to the aragonite polymorph.) Although calcite becomes thermodynamically unstable just below this, dissolution proceeds only extremely slowly. The (greater) depth at which dissolution impacts become noticeable is termed the calcite [102]. In practice this is taken as the inflection point in the trend of sedimentary CaCO3 content vs. water depth. For want of a more robust definition, a chemical lysocline is sometimes defined at X =0.8, a value which marks a distinct increase in dissolution rate [6]. Deeper still, and dissolution becomes sufficiently rapid for the dissolution flux back to the ocean to exactly balance the rain flux of calcite to the sediments. This is known as the calcite (or carbonate) compensation depth (CCD). Because in the real World the boundary in depth between sediments that have carbonate present and those in which it is completely absent is gradual rather than sharp, the CCD is operationally defined, and variously taken as the depth at which the CaCO3 content is reduced to 2 or 10 wt.%. A. Ridgwell, R.E. Zeebe / Earth and Planetary Science Letters 234 (2005) 299–315 301

a

terrestrial biota CO2

4 soils marine biota 1 reefs 3

s

deep-sea

Lysocline 2 Carbonate compensation depth (‘CCD’)

continental crust

CO b 8 2

CO2 CO2 7

5

subducting ocean plate 6

Fig. 1. The global biogeochemical cycling of calcium carbonate. (a) Modes of CaCO3 transformation and recycling within the surficial system and loss to the geological reservoir (labeled d1T through d4T). #1 Precipitation of calcite by and in the open ocean; 2+ À Ca +2HCO3 YCaCO3 +CO2(aq) +H2O. #2 Carbonate reaching deep-sea sediments will dissolve during early diagenesis if the bottom water is under-saturated and/or the organic matter flux to the sediments is sufficiently high. #3 Precipitation of CaCO3 by and shelly , with a significant fraction as the aragonite polymorph. Because modern surface are over-saturated relatively little of this carbonate dissolves in situ, and instead contributes to the formation of reefal structures or is exported to the adjoining continental slopes. #4 Precipitation of CaCO3 results in higher pCO2 at the surface, driving a net transfer of CO2 from the ocean to the atmosphere. (b) Modes of CaCO3 transformation and recycling within the geologic reservoirs and return to the surficial system (labeled d5T through d8T). #5 CaCO3 laid down in shallow as platform and reef and can be uplifted and exposed to through rifting and mountain-building episodes. CaCO3 can then 2+ À be directly recycled; CO2 +H2O+CaCO3YCa +2HCO3 . #6 Thermal breakdown of carbonates subducted into the mantle or deeply buried. The decarbonation reaction involved is essentially the reverse of , and results in the creation of calcium and release of 2+ À CO2; CaCO3 +SiO2YCO2 +CaSiO3. #7 Weathering of silicate rocks; 2CO2 +H2O+CaSiO3YCa +2HCO3 +SiO2. #8 Emission to the atmosphere of CO2 produced through decarbonation. This closes the carbon cycle on the very longest time-scales. 302 A. Ridgwell, R.E. Zeebe / Earth and Planetary Science Letters 234 (2005) 299–315

1 carbon in its inorganic, calcium carbonate (CaCO3) only minor importance globally, and will not be form also affects atmospheric CO2, but by more subtle discussed further here. means. It also plays a fundamental role in regulating While not in itself sufficient to drive substantial ocean chemistry and pH—a major factor in the abiotic precipitation, the saturation state of the modern viability of calcareous marine organisms. ocean surface is favorable to the preservation of Before exploring some of the roles that the global carbonates deposited in shallow water (neritic) envi- carbonate cycle plays in the functioning of the Earth ronments. Long-term accumulation of this material system (Section 2), we first discuss the two constit- can result in the formation of extensive marine uent parts of this cycle; (a) precipitation and burial of topographical features such as barrier reefs and CaCO3, and (b) weathering and geologic recycling, carbonate banks and platforms. A different fate awaits illustrated in Fig. 1a and b, respectively. Then, in CaCO3 precipitated in the open ocean by plankton Sections 3 and 4 we highlight the ways in which such as coccolithophores and foraminifera, however. carbonate cycling on Earth has evolved through time, This is because oceanic waters become increasingly and look to the future and the increasing impact that less saturated with depth. Below the depth of the fossil fuel CO2 release will have on the system. We saturation horizon, conditions become under-saturated finish with a brief perspective on the implications for (X b1.0) and carbonate will start to dissolve. In the future research. Readers are referred to Box 1 for a modern ocean the calcite saturation horizon (see Box brief primer on aqueous carbonate chemistry and 1) lies at about 4500 m in the Atlantic and ~3000 m in CaCO3 thermodynamics. the Pacific Ocean. Within a further 1000 m sediments are often completely devoid of carbonate particles (the 1.1. Carbonate precipitation and burial carbonate compensation depth, or dCCDT). Topo- graphic highs on the ocean floor such as the mid- Today, the surface of the ocean is everywhere Atlantic ridge can thus be picked out by sediments more than saturated (dover-saturatedT) with respect to rich in CaCO3 while the adjacent deep basins are low the carbonate phase, with a mean value for the in %CaCO3 (Fig. 2). The visual effect has been saturation state (dXT—see Box 1) of calcite of 4.8. In likened to dsnow-capped mountainsT. The pressure other words, the minimum thermodynamical require- induced surface-to-deep vertical contrast in X is ment for calcite to precipitate is exceeded by a factor further enhanced by the respiration of organic matter of almost 5 (for aragonite, X is 3.2). Despite this, the and release of metabolic CO2 in the ocean interior spontaneous precipitation of CaCO3 from the water which suppresses the ambient carbonate ion concen- column is not observed in the ocean [2]. This is tration and thus X (see Box 1). The greater because the initial step of crystal nucleation is accumulation of metabolic CO2 in the older water kinetically unfavorable, and experimentally, sponta- masses of the deep Pacific explains why the sea-floor neous (homogeneous) nucleation does not occur in there is much poorer in %CaCO3 compared to the sea water solutions until Xcalcite N~20–25 [3]. Atlantic at a similar depth [4] (Fig. 2). Although CaCO3 precipitation occurs as Unfortunately, the carbonate cycle does not con- and coatings in the marine environment, it is form to this simple picture, and a significant fraction primarily associated with the activities of living of CaCO3 appears to dissolve in the water column organisms, particularly corals, benthic shelly ani- even before it can reach the sediment surface [4–6]. mals, plankton species such as coccolithophores and This has been something of an enigma because the foraminifera, and pteropods, and where it takes place reduction in carbonate flux measured by sediment under direct metabolic control. In comparison, traps occurs well above the depth at which calcite carbonate deposition in fresh-water systems is of becomes thermodynamically unstable. Dissolution of carbonate particles in acidic digestive conditions of zooplankton guts has been one proposed mechanism 1 For the purposes of this review we simply refer to dcalcium carbonateT, but recognize that carbonates exhibit a range of [6]. Acidic micro-environments within individual substitutions of Ca2+ by Mg2+ with a generic composition of dmarine snowT aggregates may also be important

Mgxd Cad (1Àx)d CO3. [7]. Another possible explanation surrounds the A. Ridgwell, R.E. Zeebe / Earth and Planetary Science Letters 234 (2005) 299–315 303

90 100

80 3 60 0 40 wt% CaCO 20

0 -90 0 90E 180E 270E 0

Fig. 2. Distribution of the calcium carbonate content of the surface sediments of the deep sea [10]. There is an apparent predominance of CaCO3 accumulation taking place in the Atlantic and Indian Oceans compared to much more sparse accumulation in the Pacific. This is primarily a consequence of the greater accumulation of metabolic CO2 in deep Pacific waters which drives a greater degree of under-saturation and lowers the depth of the lysocline (see Box 1). The virtual absence of CaCO3 in sediments of the Southern Ocean is due to a combination of much lower CaCO3/POC rain ratio to the sediments and relatively corrosive bottom-waters. Topographic dhighsT can be picked out as areas of higher wt.% CaCO3 compared to sediments elsewhere in the same basin at similar latitudes. Areas with no data coverage (parts of the Southern Ocean, and many of the continental margins) are left blank. aragonite polymorph because it becomes susceptible 1.2. Weathering and carbonate recycling to dissolution at much shallower depths than calcite under the same ambient conditions. In support of this The weathering of calcium carbonate and calcium are recent estimates of the depth at which most silicate in soils and at exposed surfaces CaCO3 dissolution occurs in the ocean which helps balance the CaCO3 loss by appears to correspond to the aragonite saturation unlocking Ca2+ from the geologic reservoir. Alteration horizon [4,5]. However, calculations suggest that of ocean crust by percolating fluids adds an additional solute release from sinking pteropod shells, the main but more minor contribution [13]. The weathering aragonite product in the open ocean, should mostly reactions (see Fig. 1b; #5 and #7) provide the other occur much deeper than this [8]. Dissolution of raw material necessary for carbonate precipitation— À aragonite also does not help explain how 65% of bicarbonate ions (HCO3 ). However, because the À calcitic foraminiferal tests can be lost at shallow transformation 2CO2 Y 2HCO3 (Fig. 1b; #7) is depths [9]. This uncertainty is of concern because a internal to the surficial system and does not represent full appreciation of the controls of atmospheric CO2 a source of dnewT carbon, the component of CaCO3 and response to global change requires an under- burial derived from silicate rock weathering represents standing of the dissolution and the depth of recycling a loss of carbon to the geologic reservoir. This must be of CaCO3 in the water column. replaced on the long-term, achieved through the Overall, more than 80% of all carbonate precipi- release of CO2 to the atmosphere from volcanic tated in the open ocean dissolves either in the water sources2 [1]. (In contrast, the weathering and burial column or within the uppermost layers of the under- of CaCO3 results in no net loss or gain of CO2 to the lying sediments [4,10,11]. The remainder, some 1 Gt surficial system). À1 CaCO3 yr accumulates in deep-sea sediments. This burial loss, to which can be added as much as another 1 À1 2 Gt CaCO3 yr of deposition in neritic environments Imbalances between the rates of burial of organic carbon and (although the uncertainty in this figure is substantial) weathering of ancient organic matter () exposed at the land surface affects the inventory of carbon in the surficial reservoirs and [11,12], must somehow be balanced if the ocean is not thus atmospheric CO2. The details of how this particular sub-cycle to run out of calcium ions! This is achieved through the fits into the Earth system picture lies outside the scope of this weathering of carbonate and silicate rocks. review, however. 304 A. Ridgwell, R.E. Zeebe / Earth and Planetary Science Letters 234 (2005) 299–315

A powerful regulatory mechanism of the Earth glacial minimum of 189 ppm to 265 ppm at the system arises because weathering rates respond to beginning of the [15]. surface temperature and atmospheric CO2 while The demise of the great Northern Hemisphere ice simultaneously silicate weathering rates control the sheets was marked by a rise in sea-level of about À rate of transformation CO2YHCO3 and thus rate of 120 m [16]. With the flooding of the continental loss of carbon through CaCO3 burial. This is a negative shelves came a 4-fold increase in the area of shallow feedback system [14] and acts to regulate the concen- water environments available for growth [17]. tration of CO2 in the atmosphere over hundreds of Because an increase in the rate of CaCO3 deposition thousands of years [1]. will drive more CO2 into the atmosphere, this Buried carbonate is eventually recycled back from mechanism was once proposed as an explanation the geologic reservoir. This can occur if carbonates laid for the 70–80 ppm deglacial rise in atmospheric down in shallow seas such as or chalks are CO2—known as the dcoral reefT hypothesis [18]. subsequently uplifted and exposed to weathering as a Subsequent ice core measurements made it apparent result of mountain building episodes. However, carbo- that the main increase in CO2 occurred prior to the nates deposited in open ocean sediments are only rise in sea-level [19]. However, one cannot reject a infrequently exposed at the Earth’s surface, as ophiolite role for corals out of hand because reconstructions of complexes—portions of the oceanic crust and over- the time-history of reef building episodes are lying sediments that have been trapped between unambiguous in demonstrating a profound increase colliding cratonic blocks and uplifted. Instead, the in CaCO3 deposition following the end of the last primary recycling of deep-sea CaCO3 occurs through glacial [20–22]. A priori geochemical reasoning into the upper mantle and decarbonation argues that this must translate into a net re-partition- (see Fig. 1b; #6). ing of CO2 from the ocean to atmosphere. The At this point it is important to recognize that solution to this is that the dcoral reefT mechanism is carbonate burial represents the principal geologic essentially a Holocene phenomenon, with post- mechanism of CO2 removal from the ocean and glacial coral re-colonization and reefal buildup atmosphere. However, the act of precipitating CaCO3 potentially explaining much of the 20-ppm increase has the effect of re-partitioning dissolved carbon in in CO2 observed in ice cores that starts at around 8 the surface ocean into CO2(aq), raising ambient pCO2 ka [23]. and pH (see Box 1). Thus, precipitation and The global carbonate cycle plays other interesting deposition of CaCO3 have the short-term effect of biogeochemical games. Since the last glacial, the increasing the concentration of CO2 in the atmos- expansion of ecosystems to higher latitudes and phere at the expense of the ocean carbon inventory, stimulation of productivity by rising atmospheric but at the same time represents the ultimate long-term CO2 resulted in an increase in the amount of for CO2. contained in the terrestrial biosphere (vegetation plus soils). The estimates for this increase vary—from around 600 Gt C based on deep-ocean 13C changes 2. The role of the global carbonate cycle in the [24] (but see [25] for a new re-assessment), ~850 Gt C Earth system according to global vegetation models [26], to 1300 Gt C (and higher) in some paleo vegetation reconstruc- Over millions of years the silicate rock weath- tions [27]. A transfer of carbon into the terrestrial ering feedback controls the concentration of CO2 in biosphere of just 500 Gt C should have driven the atmosphere [1,14]. On time-scales shorter than atmospheric CO2 downwards by some 40 ppm [28], ca. 100 ky, however, the weathering feedback is yet ice cores show an increase between glacial and ineffective and the marine carbonate cycle plays an early Holocene of 70–80 ppm [15]. However, as CO2 important role in determining atmospheric CO2.We is sucked out of the atmosphere and ocean, oceanic 2À illustrate this by considering some of the global CO3 concentrations (and pH) increase (Fig. 3) changes that marked the end of the last ice age 18 enhancing the stability of CaCO3 in deep-sea sedi- thousand years ago (18 ka), when CO2 rose from a ments. Increased carbonate burial drives more CO2 A. Ridgwell, R.E. Zeebe / Earth and Planetary Science Letters 234 (2005) 299–315 305

-1.5 H+ OH- -2.0 )] -1 - 2- -2.5 CO2(aq) HCO3 CO3

-3.0 modern sea-water

-3.5

[concentration (mol kg -4.0 10

Log -4.5

-5.0 0 2 4 6 8 10 12 14 pH

Fig. 3. The concentrations of the dissolved carbonate species as a function of pH (referred to as the Bjerrum plot, cf. [44]): Dissolved carbon À 2À + À dioxide (CO2(aq)), bicarbonate (HCO3 ), carbonate ion (CO3 ), hydrogen ion (H ), and hydroxyl ion (OH ). At modern pH, most of À 2À the dissolved inorganic carbon is in the form of bicarbonate. Note that in seawater, the relative proportions of CO2, HCO3 , and CO3 control the pH and not vice versa.

into the atmosphere, countering about 60% of the waters where CaCO3 dissolution takes place is impact of re-growth in the terrestrial biosphere to leave determined not only by X of the overlying waters a net CO2 fall of just 17 (rather than 40) ppm [28]. This but also by the amount of metabolic CO2 released amelioration of a perturbation of atmospheric CO2 by by the in situ respiration of particulate organic changes induced in the preservation of CaCO3 in deep- carbon (POC) [34,35]. Any change in the POC flux sea sediments is known as dcarbonate compensationT to the sediments will therefore alter the fraction of [29] and represents a critical regulatory mechanism in CaCO3 that dissolves. (Strictly, it is the ratio the modern global carbon cycle on time-scales of 5–10 between CaCO3 and POC fluxes, the CaCO3/POC ky. (Carbonate compensation also represents an addi- drain ratioT that is the critical parameter rather than tional way of helping to explain the 20 ppm late the absolute POC or CaCO3 flux, per se). Models Holocene rise in atmospheric CO2 [30], to which a rise predict an atmospheric CO2 sensitivity of about 1.6 in sea-surface temperature (SST) [31] and a reduction ppm per percent reduction in CaCO3/POC [36,37]. in terrestrial carbon storage could also have contrib- A 67% increase in pelagic POC production (or 40% uted [32,33].) Carbonate compensation on its own decrease in CaCO3) could therefore theoretically does not explain why atmospheric CO2 should have account for the entire deglacial CO2 rise. Thus, risen during deglaciation at about the same time as the although the responsiveness of deep-sea sedimentary terrestrial biosphere was accumulating carbon. Clearly, CaCO3 preservation offers a means of stabilizing there must be additional carbon cycle mechanisms ocean chemistry through carbonate compensation, operating at this time to explain the ice core CO2 the atmospheric CO2 control setting on this carbo- record, the main candidates being: higher SSTs, nate regulator can be adjusted by changing surface reduced sea-ice cover, a more restricted supply ocean productivity and CaCO3/POC rain ratio. to the ocean biota, and increased ventilation of the However, despite its potential for explaining the deep ocean [23,28]. ice core CO2 record, a primary role for the rain ratio Yet another possible way of explaining an mechanism does not appear consistent with recon- increase in atmospheric CO2 arises because the structed shifts in the CCD and lysocline and model saturation state of the deep-sea sedimentary pore- analysis [38,39]. Recent interpretations of sediment 306 A. Ridgwell, R.E. Zeebe / Earth and Planetary Science Letters 234 (2005) 299–315 trap data also question whether changes in the the dated carbonate record which extends back to at CaCO3/POC rain ratio at the surface would be least 3800 Ma [48] and deposition of the first facies transmitted to the abyssal sediments [40,41] (see would have occurred well before this. Section 4). Early carbonates are characterized by A different facet of the carbonate cycle is in its sea-floor encrustations, crystal fans, and thick role as regulator of the saturation state of the ocean. beds, all indicative of a relatively rapid and dabioticT Understanding past changes in surface saturation (X) mechanism of CaCO3 precipitation. Progressively provides the environmental context for the geological younger Precambrian rocks show a decreasing interpretation of primary carbonate mineralogy, abundance of such inorganically precipitated carbo- particularly the occurrence of abundant environ- nates [49]. This secular trend in carbonate fabric mentally controlled carbonates such as marine most likely reflects a progressive decline in the cements and [42]. The occurrence of extremes degree of ocean over-saturation. However, the in X may also be important in understanding the reasons for this are not entirely clear. One possibility evolutionary driving force behind the advent of is because as the atmosphere and surface ocean biomineralizing species [43]. Furthermore, given become more oxygenated towards the end of the 2+ the partial pressure of CO2 ( pCO2), knowledge of Precambrian, sea-water concentrations of Mn and X (plus temperature and major cation composition) Fe2+ would have declined [49,50].Sincethese uniquely determines the state of the entire aqueous cations inhibit the precipitation of CaCO3, a reduc- carbonate system [44]. Thus, as -based recon- tion in their concentration would mean that a lower structions of paleo atmospheric CO2 for the Phaner- degree of over-saturation is required to achieve the ozoic improve [45] an understanding of how ocean same global carbonate deposition rate. Alternatively, X has also changed through time would enable all the gradual accretion of continental crust and the properties of the aqueous carbonate state to be associated increase in area of shallow water deposi- deduced, providing critical information in the inter- tional environments means that a lower precipitation pretation of Earth history [46]. rate per unit area (and thus X) would be required to balance the same global weathering flux [51]. What- ever the reasons, the Precambrian inorganic geo- 3. Evolution of the global carbonate cycle through chemical age was brought to a relatively abrupt end Earth history as life stepped on the evolutionary accelerator and drove the Earth system through a succession of new We review the history of global carbonate cycling modes of carbonate cycling. in two parts; the Precambrian (up to 542 Ma), when inorganic geochemical processes tended to dominate 3.2. Carbonate cycling in the Phanerozoic—enter the the nature and location of carbonate deposition, and biota the Phanerozoic (542 Ma to present), when life became the single most important factor. The advent of carbonate occurred around the time of the –Precambrian boun- 3.1. Carbonate cycling in the Precambrian—when dary [52] when evolutionary innovation conferred on ruled the roost organisms the ability to precipitate carbonate struc- tures (). Prior to this there could have been no The requirements for carbonate cycling to begin on significant biologically driven production of CaCO3, the early Earth are fairly minimal—the contact of and carbonate deposition would have been primarily basaltic rock with water and dissolved CO2 to initiate restricted to heterogeneous nucleation and crystal chemical weathering [47]. With the weathering of growth on organic and inorganic surfaces in warm silicate rocks comes the delivery of solutes to the shallow water environments [49]. Because biominer- ocean, making an over-saturated surface and the alization enabled the more efficient removal of weath- eventual precipitation of carbonates inevitable. The ering products from the ocean by the expenditure of early start to this is reflected in metabolic energy, a lower thermodynamic driving A. Ridgwell, R.E. Zeebe / Earth and Planetary Science Letters 234 (2005) 299–315 307 force for carbonate precipitation would have been highly over-saturated ocean. This is consistent with the required in the ambient marine environment. The widespread occurrence of abundant environmentally result would have been a reduction in ocean saturation controlled carbonates such as cements, calcified (X) as the Phanerozoic got under way. , and thick precipitated beds during parts A second major development took place several of the and [42,65,66], all indicative of hundred million years (My) later, with the Mesozoic comparatively rapid and dabioticT modes of carbonate proliferation of planktic calcifiers [53] and the precipitation. However, despite similar sea-level and establishment of the modern mode of carbonate cation chemistry, environmentally controlled carbo- cycling in a dMid-Mesozoic RevolutionT [46].We nates are rare in the modern ocean. The difference is a illustrate the profound importance of this by consid- direct consequence of the proliferation of calcareous ering the response of the marine carbonate cycle to two plankton during the Mesozoic and creation of a new environmental forcings; (i) sea-level, which varies and substantive sink for CaCO3 [46,65,66]. over hundreds of million years by up to 300 m (Fig. Although benthic foraminifera and other bottom- 4a), and (ii) the cation chemistry of the ocean; dwelling calcifiers evolved early in the Phanerozoic, it particularly (Mg2+) and Calcium (Ca2+) is not until the Mesozoic that a marked proliferation in ion concentrations (Fig. 4c). Although global temper- and planktic foraminiferal diversity atures and continental paleo-latitude also affect global and abundance is observed [53,67] (Fig. 4b). Only carbonate deposition by determining the latitudinal then would a substantive deep-sea sedimentary carbo- extent of carbonate production by warm-water corals nate sink have been possible. This supposition is [43,54], we will restrict our analysis to just two factors. supported by the observed composition of Phanerozoic Times of high sea-level such as the Mid ophiolite suites which indicate that pelagic carbonate produced flooding extents in excess of 50% on some accumulation was comparatively rare in Paleozoic cratons [55] and the creation of extensive inland ocean sediments [68] (Fig. 4e). Conversely, the mean (epeiric) seas. This in turn facilitated widespread area of platform carbonates during the Mesozoic and carbonate platform development and shallow water Cenozoic is much reduced compared to the Paleozoic carbonate accumulation [43,54] which would have (Fig. 4f). One might speculate whether the ca. 200 My driven a lower X. Conversely, times of low sea-level gap between the first appearance of calcifying planktic and restricted depositional area would produce a foraminifera and coccolithophorids and their rise to tendency towards high ocean X [56]. Superimposed relative dominance in global pelagic ecosystems [53] on this is a variation in the oceanic ratio of Mg2+ to is related to extreme ocean over-saturation in the late Ca2+ by a factor of three [57–59]. In order to maintain Permian and early Triassic, a potential environmental the same global rate of carbonate production, higher driving force favoring calcifiers. A similar thesis can ambient Mg2+/Ca2+ requires a more over-saturated be advanced to help explain the timing of the advent of ocean because Mg2+ inhibits calcite precipitation [60]. metazoan biomineralization following the inferred (The inhibition is predominantly a result of the Mg2+ occurrence of extreme oceanic saturation events interaction with the solid calcite phase, rather than a during the late Precambrian [69]. 2+ 2À solution effect involving Mg –CO3 complexation. The establishment of a substantive deep-sea 2À The latter effect reduces CO3 activity but is carbonate sink is important because it introduced a independent of the CaCO3 polymorph present in new stabilizing mechanism to the Earth system— solution—see [61] and references therein.) At higher dcarbonate compensationT (see Section 2). Indeed, the X, aragonite becomes more common in abiotic absence of a responsive deep-sea carbonate sink in the cements and hyper-calcifying organisms, which we Precambrian would have made the carbon–climate observe in the geological record as distinctive times system much more sensitive to perturbation. Ice ages of relatively abundant shallow water carbonate of near-global extent and multi million-year duration aragonite [62,63]—periods dubbed daragonite seasT deduced for the end of the Precambrian [70] could [64] (Fig. 4d). have been facilitated by the weak dbufferingT of the The coincidence of times of low sea-level and low Precambrian carbon cycle and atmospheric CO2 Ca2+ concentrations (Fig. 4) should have given rise to a [71,72]. This view is also consistent with the wide- 308 A. Ridgwell, R.E. Zeebe / Earth and Planetary Science Letters 234 (2005) 299–315

spread occurrence of strange dcapT carbonate facies high CO2 atmosphere [73,74] and the overturning of a deposited during postglacial flooding of the shelves. stagnant ocean [42]. However, all hypotheses recog- Other explanations for the genesis of cap carbonates nize extreme changes taking place in global carbonate in the aftermath of extreme late Precambrian glacia- cycling at this time. tion have been proposed, such as the removal of the As well as adding new mechanisms for stabilizing solutes derived from rapid rock weathering under a atmospheric CO2, the establishment of a substantive

CENOZOIC MESOZOIC PALEOZOIC PRE-CAMBRIAN

N Pg CJTrPrCDSO ∈ E

300 200 a 100 0 sea-level (m) -100

Dinoflagellates Acritarchs Coccolithophorids

Major changes in Planktic foraminifers plankton assemblage b 60 ) -1

40

] (mmol kg 20 some of the forcings global carbonate cycle 2+ c [Ca 0

calcite dominantaragonite dominant calcite dominant (low Mg2+/Ca2+) (high Mg2+/Ca2+) (low Mg2+/Ca2+) d 100 deep-sea deposition dominant 80 60 40

% occurrence 20 e )

2 0 km 3

6 neritic CaCO deposition dominant 40 3 30 20

10 as recorded in the geological record response of global carbonate cycling

area of CaCO f 0

accumulation (10 0 100 200300 400 500 600 Age (Ma) A. Ridgwell, R.E. Zeebe / Earth and Planetary Science Letters 234 (2005) 299–315 309

deep-sea sedimentary CaCO3 sink would have also had changes in the population or ecological success of a destabilizing effect. For instance, episodes of high shallow-water calcifiers. The Mesozoic shift towards weathering rates and sequestration of carbon in pelagic widespread pelagic biomineralization finally led to a carbonates could subsequently to periods of significant stabilization of the marine CaCO3 satu- enhanced metamorphic CO2 out-gassing to the atmos- ration state, termed the dCretanT ocean. Large and phere as the sea-floor CaCO3 is subducted and under- rapid shifts between, e.g. the Neritic- and Cretan- goes decarbonation [75]. The subduction of carbonates ocean mode have likely occurred in the aftermaths of as ocean basins close and are destroyed would also catastrophic events such as the –Tertiary result in episodic enhanced CO2 release [76,77]. Both bolide impact [78]. mechanisms predict secular oscillation in metamorphic CO2 out-gassing rates on tectonic time-scales, and both would not have been possible before the Mid- 4. Back to the future: carbon cycling in the Mesozoic Revolution in carbonate deposition. Anthropocene

3.3. Synthesis The ocean is capable of absorbing about 70% of all the CO2 released by fossil fuel combustion [79]. For a This evolution in global carbonate cycling over 4000 Gt C dburnT, this means that the equivalent of geologic time can be neatly conceptualized as three ~600 ppm CO2 will remain in the atmosphere after distinct marine carbonate cycle modes, termed hundreds of years [80]. An atmospheric CO2 concen- dStrangeloveT, dNeritanT, and dCretanT ocean modes tration of ~1000 ppm is about three times the present- by Zeebe and Westbroek [39]. The geochemistry- day (year 2003) value of 376 ppm [81] and represents ruled Precambrian mode of CaCO3 cycling resembles a very significant long-term radiative forcing of the a carbonate-dStrangeloveT ocean, in which biogenic climate system. The terrestrial biosphere is unlikely to precipitation of CaCO3 is essentially absent. It is be of much help and may well become a net source of characterized by high-supersaturation and generally CO2 to the atmosphere in the coming centuries as the inorganic (at most partly biologically mediated) Earth’s surface warms [82,83], further exacerbating formation of carbonates. Following the advent of the problem. Fortunately, geochemical interactions biomineralization in the Cambrian, biologically con- between the ocean and deep-sea sediments intervene trolled carbonate precipitation in shallow-water (ner- and on a time-scale of 5–8 ky, carbonate compensa- itic) environments became significant. Its conceptual tion will remove a further 10–20% of fossil fuel CO2 analog is the dNeritanT ocean, in which the dominant emitted to the atmosphere [79]. Ultimately, on a time- 2+ 2À mode of Ca and CO3 removal from seawater is scale of 1 My, higher silicate rock weathering rates biogenic, neritic carbonate deposition. The saturation induced by enhanced greenhouse warming will state of the Neritan ocean is highly susceptible to remove the remaining fraction.

Fig. 4. Evolution of global carbonate cycling through the Phanerozoic—major driving forces (panels a though c) and responses of the system recorded in the geological record (panels d through f). FORCINGS: (a) Eustatic sea-level plotted relative to modern [103]. (b) Major changes in plankton assemblages [53]. Calcifying taxa are highlighted in black with non-calcifying taxa shown in grey. Although the rise of planktic foraminiferal taxa (for which we take Globigerinina as broadly representative) occurs during the early- to mid-Mesozoic, the evolution of the first calcifying foraminifera taxa occurred somewhat earlier in the mid-Paleozoic [53]. (c) Paleo-marine Ca2+ concentrations as recorded in fluid inclusions contained in marine halite crystals [57] (vertical grey bars) as well as the model predictions of [62] (black curve). RESPONSES: (d) Ascribed characteristic periods of aragonite and calcite dseasT [62,63], corresponding to times of high and low Mg2+/Ca2+ ratios, respectively. (e) Percent occurrence of carbonates in ophiolite complexes for which sedimentary composition has been reported [68]. The comparative rarity of carbonate sediments in ophiolite complexes of Paleozoic age is noticeable. (f) Reconstructed changes in the total area of platform (shallow water) carbonates [43]. A general poleward movement of the major continents through the Phanerozoic has been proposed as the reason for the apparent long-term decline in areal extent [43]. However, the general shape of the curve is also consistent with an oscillation in sea-level, with a step reduction in the importance of shallow water carbonate after the Mid-Mesozoic Revolution [46]. The geological periods from the end- Precambrian to end-Phanerozoic are delineated at the top, running from Ediacran (dET) (600 to 542 Ma) at the far right through to Paleogene (dPgT) and Neogene (dNT) on the left-hand side. 310 A. Ridgwell, R.E. Zeebe / Earth and Planetary Science Letters 234 (2005) 299–315

This has been the view of the response of the historical fossil fuel CO2 emissions, surface pH has global carbonate cycle to anthropogenic perturba- already been reduced by some ~0.1 pH units tion—largely predictable and beneficial. New research [80,81]. A further fall over the next few hundred a much murkier picture. years of more than 0.6 pH units is possible [80] (Fig. When CO2 gas dissolves in water, there is a 5). Because carbonate is thermodynamically less reduction in carbonate ion concentrations and an stable under such conditions, the metabolic cost to increase in ocean acidity (see Box 1). As a result of organisms of building carbonate shells and skeletons

CENOZOIC MESOZOIC PALEOZOIC

N Pg CJTrPrCDSO ∈ Year PHANEROZOIC 3000 2750 2500 2250 2000 1750 7000 a 6000 5000

4000 µ

3000 2 CO ( atm)

2000 p 1000 0 8.5 b H p 8.0

7.5 mean surface

7.0 0 100 200300 400 500 Age (Ma)

Fig. 5. A geologic perspective on current anthropogenic perturbation of the global carbon cycle. (a) Phanerozoic evolution of atmospheric

CO2 reconstructed from proxy records by Royer et al. [45]. Paleo-CO2 data has been binned into 20 My intervals, with the mean and error (one standard deviation) for each interval shown as horizontal black dash and vertical grey bar, respectively. The left hand side of the figure shows the historical atmospheric CO2 trend (years 1800 to 2000) followed by the range in trajectories that would occur if between 4000 and 8000 Gt C fossil fuel resources were to be combusted (and also depending on the assumed rate of CO2 emissions) [104]. A peak value of between ~1000 and 3000 ppm is reached before the end of this millennium, indicated by the pair of horizontal dashed lines. (b) Model- predicted evolution of mean surface pH through the Phanerozoic (but only considering the case of a modern mode of carbonate cycling)

[46]. The solid black line represents the response of the global carbonate cycle to the mean paleo pCO2 reconstruction while the grey-filled 2+ envelope reflects the response to the error (1 S.D.) in paleo CO2. The model is also forced with changes in ambient ocean Ca concentrations (see Fig. 4c) following [62]), which has the effect of additionally suppressing ocean pH by up to ~0.25 pH units during periods of elevated [Ca2+] such as the early-to-mid Paleozoic, and mid-to-late Mesozoic. Further factors affecting carbonate cycling have a comparatively smaller effect and are excluded for clarity. For instance, the absence of a significant deep-sea sedimentary carbonate sink prior to ca. 200 Ma would increase pH and make the earlier Phanerozoic ocean slightly less acidic compared to the curve shown here, but by no more than ~0.1 pH units [46]. The predicted historical and future trajectory of mean surface ocean pH in response to the same range of CO2 emission scenarios as detailed in (a) above [104] is shown on the left hand side of the figure. Future surface ocean pH reaches a minimum in the range 7.7–7.25 (indicated by the pair of horizontal dashed lines). A. Ridgwell, R.E. Zeebe / Earth and Planetary Science Letters 234 (2005) 299–315 311 H) p (high 2 low CO H) p (low 2 high CO

Fig. 6. Scanning electron (SEM) photographs of coccolithophorids cultured under different CO2 concentrations [90]. (a,b,d,e) and (c,f) Gephyrocapsa Oceanica. Scale bars represent 1 Am. structure is notably different, with distinct malformations and a reduced degree of in cells grown at elevated CO2 levels (and lower pH) (d,e,f) compared to cultures incubated at preindustrial CO2 levels (a,b,c). With copyright permission from Nature Publishing Group. will be greater. Experimental studies have indeed However, a strong association between partic- demonstrated that corals are adversely affected at ulate organic carbon (POC) and CaCO3 fluxes higher pCO2 [84–86] with the implications for coral recognized in deep-sea sediment trap data [92,40] reef ecosystems already starting to be widely may reflect a dballastingT3 of organic matter by appreciated [87,88]. In contrast, much less attention carbonate particles [92,93]. A decrease in CaCO3 has been paid to the open ocean environment. Only production would then drive a reduction in the more recently has the extent to which calcifying efficiency with which POC is transported to depth, plankton could be affected started to be recognized weakening the , and driving higher [89–91] (Fig. 6). This has important implications for surface ocean pCO2 [94]. This would reduce the flux the marine carbon cycle. of fossil fuel CO2 into the ocean [95] and exacerbate The precipitation of CaCO3 and subsequent future . removal from surface ocean layers through gravita- Which is it? Unfortunately, the respective impor- tional settling have the effect of driving surface ocean tance of these two mechanisms is poorly constrained pCO2 higher. This decreases the air–sea CO2 gradient and even the sign of the net impact is uncertain [95]. and opposes the uptake of fossil fuel CO2 from the Underlying this is uncertainty in the interpretation of atmosphere. If carbonate production was reduced, the observed relationship between sinking fluxes of surface ocean pCO2 would fall and the rate of CO2 CaCO3 and POC because alternative explanations invasion into the ocean would be enhanced—acting as require no direct link between the degree of a dbrakeT (negative feedback) on rising atmospheric calcification and the efficiency of particulate organic CO2 [89–91]. Preliminary estimates suggest that an matter transport [40,41]. increase in the rate of CO2 uptake of 0.5–1.0 Gt C yrÀ1 is possible [91]. In this respect, decreasing calcification and CaCO3 export rates would play a direct and dhelpfulT role in ameliorating future global 3 The enhancement of the sinking rate of POC through the water change. column due to a greater mean aggregate . 312 A. Ridgwell, R.E. Zeebe / Earth and Planetary Science Letters 234 (2005) 299–315

5. Summary and perspectives water column (and the general controls on the CaCO3/ POC drainT ratio at the sediment surface), and (iv) The trajectory that the concentration of CO2 in the application of coupled carbon-climate models in the atmosphere takes will largely dictate the rate and quantitative interpretation of dcatastrophicT geological magnitude of future climate change. In order to make events such as the PETM. sufficiently informed choices regarding the maximum fossil fuel CO2 release that will keep global change Acknowledgements within dacceptableT limits, the natural pathways of CO2 removal from the atmosphere must be fully charac- terized. Of primary importance is the oceanic sink, AJR acknowledges support from Canada Research which already accounts for the equivalent of almost Chairs. Both authors would like to thank Derek Vance, Daniela Schmidt, Brad Opdyke, and an half of all CO2 emissions due to fossil fuel burning and cement manufacture [96]. Improving our understand- anonymous reviewer for helpful and supportive com- ing of the role of the marine carbonate cycle in this and ments on the manuscript. how it might change in the future is essential. To put the possible environmental changes facing us into some perspective, one would have to turn the References clock back at least 100 million years to find analogous surface ocean pH conditions (Fig. 5b). One must [1] R.A. Berner, K. Caldeira, The need for mass balance and recognize, however, that the calcifying species feedback in the geochemical carbon cycle, 25 (1997) 955–956. involved and relative importance of shallow vs. [2] J.W. Morse, S.L. He, Influences of T, S and pCO2 on the deep-water carbonate deposition have both changed pseudo-homogeneous precipitation of CaCO3 from sea- over this period. More importantly, this also excludes water—implications for whiting formation, Mar. Chem. 41 transient perturbations of the carbon cycle— (1993) 291–297. dcatastrophicT events such as associated with the [3] J.W. Morse, D.K. Gledhill, F.J. Millero, CaCO3 precipitation d T kinetics in waters from the Great Bahama Bank: implications /Eocene Thermal Maximum (PETM) at for the relationship between Bank, Geochim. Cosmochim. 55.5 Ma [97]. Indeed, the early Eocene could have Acta 67 (2003) 2819–2826. seen significant surface in [4] R.A. Feely, et al., Impact of anthropogenic CO2 on the CaCO system in the oceans, Science 305 (2004) 362–366. response to inferred CO2 release to the ocean and 3 atmosphere. Events such as the PETM represent a [5] S.-N. Chung, et al., Calcium carbonate budget in the Atlantic column inorganic carbon chemistry, Glob. Biogeochem. possible geologic analogue for future global change Cycles 17 (2003), doi:10.1029/2002GB002001. [97,98]. Understanding the PETM may prove critical [6] J.D. Milliman, et al., Biologically mediated dissolution of in being able to correctly predict the long-term impact calcium carbonate above the chemical lysocline?, Deep-Sea Res., Part I 46 (1999) 1653–1669. of continued fossil fuel CO2 release, as well as what (if any) species of marine calcifying organisms might [7] H. Jansen, R.E. Zeebe, D.A. Wolf-Gladrow, Modeling the dissolution of settling CaCO3 in the ocean, Glob. go extinct. However, we are still far from achieving Biogeochem. Cycles 16 (2002), doi:10.1029/2000GB001279. this understanding and even the source and magnitude [8] R.H. Byrne, et al., Water column dissolution of aragonite in of the carbon release is currently hotly debated [99]. the Pacific Ocean, Nature 312 (1984) 321–326. Given the importance of managing future global [9] R. Schiebel, Planktic foraminiferal sedimentation and the change, we believe that new impetus should be given marine calcite budget, Glob. Biogeochem. Cycles 16 (2002), doi:10.1029/2001GB001459. to elucidating the role of the global carbonate cycle in [10] D. Archer, An atlas of the distribution of calcium carbonate the regulation and evolution of the Earth system. We in sediments of the deep sea, Glob. Biogeochem. Cycles 10 identify several key priorities; (i) better quantification (1996) 159–174. of the modern global carbonate budget, especially of [11] J.D. Milliman, A.W. Droxler, Neritic and pelagic carbonate neritic deposition, (ii) elucidating the response of sedimentation in the marine environment: ignorance is not bliss, Geol. Rundsch. 85 (1996) 496–504. planktic calcifiers and ecosystem composition to [12] A. Vecsei, A new estimate of global reefal carbonate depressed ambient pH, (iii) understanding the reasons production including the fore-reefs, Glob. . Change for the dissolution of carbonate particles sinking in the 43 (2004) 1–18. A. Ridgwell, R.E. Zeebe / Earth and Planetary Science Letters 234 (2005) 299–315 313

[13] K. Caldeira, Long-term control of atmospheric carbon- [31] F. Joos, et al., Transient simulations of Holocene atmospheric dioxide—low-temperature sea-floor alteration or terrestrial carbon dioxide and terrestrial carbon since the Last Glacial silicate-rock weathering, Am. J. Sci. 295 (1995) 1077–1114. Maximum, Glob. Biogeochem. Cycles 18 (2004), doi:10.1029/ [14] R.A. Berner, A new look at the long-term carbon cycle, GSA 2003GB002156. Today 9 (1999) 1–6. [32] V. Brovkin, et al., Carbon cycle, carbon cycle, vegetation and

[15] E. Monnin, et al., Atmospheric CO2 concentrations over the climate dynamics in the Holocene: experiments with the last glacial termination, Science 291 (2001) 112–114. CLIMBER-2 Model, Glob. Biogeochem. Cycles 16 (2002), [16] R.G. Fairbanks, A 17,000-year glacio-eustatic sea level doi:10.1029/2001GB001662. record: influence of glacial melting rates on the Younger [33] A. Indermqhle, et al., Holocene carbon-cycle dynamics based

Dryas event and deep-ocean circulation, Nature 342 (1989) on CO2 trapped in ice at Taylor Dome, Antarctica, Nature 637–642. 398 (1999) 121–126. [17] J.A. Kleypas, Modeled estimates of global reef habitat and [34] D. Archer, J.L. Morford, S.R. Emerson, A model of suboxic carbonate production since the last glacial maximum, sedimentary diagenesis suitable for automatic tuning and 12 (1997) 533–545. gridded global domains, Glob. Biogeochem. Cycles 16 [18] W.H. Berger, Increase of carbon dioxide in the atmosphere (2002), doi:10.1029/2000GB001288. during deglaciation: the hypothesis, Naturwissen- [35] B. Hales, Respiration, dissolution, and the lysocline, Paleo- schaften 69 (1982) 87–88. ceaongraphy 18 (2003), doi:10.1029/2003PA000915. [19] W.S. Broecker, G.M. Henderson, The sequence of events [36] D. Archer, E. Maier-Reimer, Effect of deep-sea sedimentary

surrounding Termination II and their implications for the calcite preservation on atmospheric CO2 concentration, cause of glacial–interglacial CO2 changes, Paleoceanography Nature 367 (1994) 260–263. 13 (1998) 352–364. [37] A.J. Ridgwell, A.J. Watson, D.E. Archer, Modelling the [20] L. Montaggioni, Postglacial reef growth, Earth Planet. Sci. response of the oceanic Si inventory to perturbation, and

331 (2000) 319–330. consequences for atmospheric CO2, Glob. Biogeochem. [21] D.A. Ryan, B.N. Opdyke, J.S. Jell, Holocene sediments of Cycles 16 (2002) 1071, doi:10.1029/2002GB001877. Wistari Reef: towards a global quantification of coral reef [38] D.M. Sigman, D.C. McCorkle, W.R. Martin, The calcite related neritic sedimentation in the Holocene, Palaeogeogr. lysocline as a constraint on glacial/interglacial low-latitude Palaeoclimatol. Palaeoecol. 175 (2001) 173–184. production changes, Glob. Biogeochem. Cycles 12 (1998)

[22] A. Vecsei, W.H. Berger, Increase of atmospheric CO2 during 409–427. deglaciation: constraints on the coral reef hypothesis from [39] R.E. Zeebe, P. Westbroek, A simple model for the CaCO3 patterns of deposition, Glob. Biogeochem. Cycles 18 (2004), saturation state of the ocean: the bStrangelove,Q the bNeritan,Q doi:10.1029/2003GB002147. and the bCretanQ Ocean, Geochem. Geophys. Geosyst. 4 [23] A.J. Ridgwell, A.J. Watson, M.A. Maslin, J.O. Kaplan, (2003), doi:10.1029/2003GC000538. Implications of coral reef buildup for the controls on [40] R. Francois, et al., Glob. Biogeochem. Cycles 16 (2002),

atmospheric CO2 since the Last Glacial Maximum, Paleo- doi:10.1029/2001GB001722. ceanography 18 (2003), doi:10.1029/2003PA000893. [41] U. Passow, Switching perspectives: do mineral fluxes [24] T.J. Crowley, Ice-age terrestrial carbon revisited, Glob. determine particulate organic carbon fluxes or vice versa?, Biogeochem. Cycles 9 (1995) 377–389. Geochem. Geophys. Geosyst. 5 (2004), doi:10.1029/2003 [25] M.A. Maslin, E. Thomas, Balancing the deglacial global GC000670. carbon budget: the hydrate factor, Quat. Sci. Rev. 22 (2003) [42] J.P. Grotzinger, A.H. Knoll, Anomalous carbonate precip- 1729–1736. itates: is the Precambrian the key to the Permian?, Palaios 10 [26] J.O. Kaplan, I.C. Prentice, W. Knorr, P.J. Valdes, Modelling (1995) 578–596. the dynamics of terrestrial carbon storage since the Last [43] L.J. Walker, et al., Continental drift and Phanerozoic Glacial Maximum, Geophys. Res. Lett. 29 (2002) 2074, carbonate accumulation in shallow-shelf and deep-marine doi:10.1029/2002GL015230. settings, J. Geol. 110 (2002) 75–87.

[27] J.M. Adams, et al., Increases in terrestrial carbon storage [44] R.E. Zeebe, D. Wolf-Gladrow, CO2 in seawater: equilibrium, from the Last Glacial Maximum to the present, Nature 348 kinetics, isotopes, Elsevier Oceanographic Series, vol. 65, (1990) 711–714. Elsevier, New York, 2001.

[28] D. Archer, A. Winguth, D. Lea, N. Mahowald, What caused [45] D.L. Royer, et al., CO2 as a primary driver of Phanerozoic the glacial/interglacial atmospheric pCO2 cycles?, Rev. climate, GSA Today 14 (2004) 4–10. Geophys. 38 (2000) 159–189. [46] A.J. Ridgwell, A mid Mesozoic revolution in the regulation

[29] W.S. Broecker, T.-H. Peng, The role of CaCO3 compensation of ocean chemistry, Mar. Geol. (in press). in the glacial to interglacial atmospheric CO2 change, Glob. [47] N.H. Sleep, K. Zahnle, Carbon dioxide cycling and Biogeochem. Cycles 1 (1987) 15–29. implications for climate on ancient Earth, J. Geophys. Res. [30] W.S. Broecker, J. Lynch-Stirglitz, E. Clark, I. Hajdas, G. 106 (2001) 1373–1399.

Bonani, What caused the atmosphere’s CO2 content to rise [48] G. Shields, J. Veizer, Precambrian marine carbonate isotope during the last 8000 years?, Geochem. Geophys. Geosyst. 2 database: version 1.1, Geochem. Geophys. Geosyst. 3 (2002), (2001) (2001GC000177). doi:10.1029/2001GC000266. 314 A. Ridgwell, R.E. Zeebe / Earth and Planetary Science Letters 234 (2005) 299–315

[49] J.P. Grotzinger, N.P. James, Precambrian carbonates; evolu- [68] S.K. Boss, B.H. Wilkinson, Planktogenic eustatic control on tion of understanding, Carbonate and Dia- cratonic oceanic carbonate accumulation, J. Geol. 99 (1991) genesis in the Evolving Precambrian World, SEPM, 2000. 497–513. [50] J.P. Grotzinger, J.F. Kasting, New constraints on Precambrian [69] S.T. Brennan, T.K. Lowenstein, J. Horita, Seawater chem- ocean composition, J. Geol. 101 (1993) 235–243. istry and the advent of biocalcification, Geology 32 (2004) [51] D.J. Des Marais, Isotopic evolution of the biogeochemical 473–476. carbon cycle during the Precambrian, Rev. Mineral. Geo- [70] J.L. Kirschivink, Late low-latitude global glaci- chem. 43 (2001) 555–578. ation: the , in: J.W. Schopf, C. Klein (Eds.), [52] R.A. Wood, J.P. Grotzinger, J.A.D. Dickson, Proterozoic The Proterozoic Biosphere, Cambridge University Press, modular biomineralized metazoan from the Nama Group, Cambridge, 1992, pp. 51–52. Namibia, Science 296 (2002) 2383–2386. [71] A.J. Ridgwell, M.J. Kennedy, K. Caldeira, Carbonate [53] R.E. Martin, Cyclic and secular variation in deposition, climate stability, and Neoproterozoic ice ages, biomineralization—clues to the biogeochemical evolution Science 302 (2003) 859–862. of Phanerozoic oceans, Glob. Planet. Change 11 (1995) [72] A.J. Ridgwell, M.J. Kennedy, Secular changes in the 1–23. importance of neritic carbonate deposition as a control on [54] B.N. Opdyke, B.H. Wilkinson, saturation the magnitude and stability of Neoproterozoic ice ages, in: state and cratonic accumulation, Am. J. Sci. 293 G. Jenkins, et al. (Eds.), The Extreme Proterozoic: Geology, (1993) 217–234. Geochemistry, and Climate, Geophysical Monograph Ser- [55] T.J. Algeo, K.B. Seslavinsky, The Paleozoic world—con- ies, vol. 146, American Geophysical Union, Washington, tinental flooding, hypsometry, and sealevel, Am. J. Sci. 295 DC, 2004. (1995) 787–822. [73] P.F. Hoffman, A.J. Kaufman, G.P. Halverson, D.P. Schrag,

[56] T. Volk, Sensitivity of climate and atmospheric CO2 to deep- A Neoproterozoic snowball earth, Science 281 (1998) ocean and shallow-ocean carbonate burial, Nature 337 (1989) 1342–1346. 637–640. [74] P.F. Hoffman, D.P. Schrag, The snowball Earth hypothesis: [57] J. Horita, et al., Chemical evolution of seawater during the testing the limits of global change, Terra Nova 14 (2002) Phanerozoic: implications from the record of marine evap- 129–155. orates, Geochim. Cosmochim. Acta 66 (2002) 3733–3756. [75] K. Caldeira, Continental–pelagic carbonate partitioning and [58] T.K. Lowenstein, et al., Oscillations in Phanerozoic seawater the global carbonate-silicate cycle, Geology 19 (1991) chemistry: evidence from fluid inclusions, Science 294 204–206. (2002) 1086–1088. [76] J.M. Edmond, Y. Huh, Non-steady state carbonate recycling

[59] T.K. Lowenstein, et al., Secular variation in seawater and implications for the evolution of atmospheric pCO2, chemistry and the origin of basinal brines, Earth Planet. Sci. Lett. 216 (2003) 125–139.

Geology 31 (2003) 857–860. [77] D.P. Schrag, Control of atmospheric CO2 and climate [60] K.J. Davis, et al., The role of Mg2+ as an impurity in calcite through Earth history, Geochim. Cosmochim. Acta 66 growth, Science 290 (2000) 1134–1137. (2002) A688. [61] R.E. Zeebe, A. Sanyal, Comparison of two potential [78] K. Caldeira, M.R. Rampino, Aftermath of the end-cretaceous strategies of planktonic foraminifera for house building: mass extinction—possible biogeochemical stabilization of Mg2+ or H+ removal?, Geochim. Cosmochim. Acta 66 the carbon-cycle and climate, Paleoceanography 8 (1993) (2002) 1159–1169. 515–525. [62] S.M. Stanley, L.A. Hardie, Secular oscillations in the [79] D. Archer, H. Kheshgi, E. Maier-Reimer, Multiple timescales

carbonate mineralogy of reef-building and sediment-produc- for neutralization of fossil fuel CO2, Glob. Biogeochem. ing organisms driven by tectonically forced shifts in seawater Cycles 12 (1998) 259. chemistry, Palaeogeogr. Palaeoclimatol. Palaeoecol. 144 [80] K. Caldeira, M.E. Wickett, Anthropogenic carbon and ocean (1998) 3–19. pH, Nature 425 (2003) 365.

[63] S.M. Stanley, L.A. Hardie, Hypercalcification: [81] C.D. Keeling, T.P. Whorf, Atmospheric CO2 records from link and geochemistry to sedimentology, GSA sites in the SIO air sampling network, Trends: A Compen- Today 9 (1999) 1–7. dium of Data on Global Change, Carbon Dioxide Informa- [64] P.A. Sandberg, An oscillating trend in Phanerozoic non- tion Analysis Center, Oak Ridge National Laboratory, U.S. skeletal carbonate mineralogy, Nature 305 (1983) 19–22. Department of Energy, Oak Ridge, Tenn., USA, 2004, (http://

[65] R. Riding, Phanerozoic patterns of marine CaCO3 precip- cdiac.esd.ornl.gov/trends/co2/sio-mlo.htm). itation, Naturwissenschaften 80 (1993) 513–516. [82] T.M. Lenton, Land and ocean carbon cycle feedback effects [66] D.Y. Sumner, J.P. Grotzinger, Were kinetics of on global warming in a simple Earth system model, Tellus 52 calcium carbonate precipitation related to concen- (2000) 1159–1188. tration?, Geology 24 (1996) 119–122. [83] J.A. Kleypas, et al., Geochemical consequences of increased [67] M.B. Hart, The search for the origin of the planktic atmospheric carbon dioxide on coral reefs, Science 284 Foraminifera, J. Geol. Soc. 160 (2003) 341–343. (1999) 118. A. Ridgwell, R.E. Zeebe / Earth and Planetary Science Letters 234 (2005) 299–315 315

[84] C. Langdon, et al., Effect of calcium carbonate saturation [99] A.C. Kurtz, et al., Early Cenozoic decoupling of the global state on the calcification rate of an experimental coral reef, carbon and sulfur cycles, Paleoceanography 18 (2003), Glob. Biogeochem. Cycles 14 (2000) 639–654. doi:10.1029/2003PA000908.

[85] N. Leclercq, J.P. Gattuso, J. Jaubert, CO2 partial pressure [100] S.J. Zhong, A. Mucci, Calcite precipitation in seawater controls the calcification rate of a coral community, Glob. using a constant addition technique—a new overall reaction Chang. Biol. 6 (2000) 329–334. kinetic expression, Geochim. Cosmochim. Acta 57 (1993) [86] A.T. Marshall, P.L. Clode, Effect of increased calcium 1409–1417. concentration in sea water on calcification and photosyn- [101] L.M. Walter, J.W. Morse, The dissolution kinetics of shallow thesis in the scleractinian coral Galaxea fascicularis, J. Exp. marine carbonates in seawater: a laboratory study, Geochim. Biol. 205 (2002) 2107–2113. Cosmochim. Acta 49 (1985) 1503–1513.

[87] T.P. Hughes, et al., Climate change, human impacts, and the [102] W.S. Broecker, The oceanic CaCO3 cycle, in: H.D. Holland, resilience of coral reefs, Science 301 (2003) 929–933. K.K. Turekian (Eds.), Treatise on Geochemistry, Elsevier, [88] J.A. Kleypas, R.W. Buddemeier, J.-P. Gattuso, Int. J. Earth 2003, pp. 529–549. Sci. 90 (2001) 426. [103] B.U. Haq, et al., Mesozoic and Cenozoic chronostratigraphy [89] S. Barker, H. Elderfield, Foraminiferal calcification response and cycles of sea-level change, in: C.K. Wilgus, et al. (Eds.),

to glacial–interglacial changes in atmospheric CO2, Science Sealevel-changes; An Integrated Approach, Special Publica- 297 (2002) 833. tion-Society of Economic Paleontologists and Mineralogists, [90] U. Riebesell, I. Zondervan, B. Rost, P.D. Tortell, R.E. Zeebe, vol. 42, 1988, pp. 71–108. F.M.M. Morel, Reduced calcification of marine plankton in [104] T. Tyrrell, Submission to the Royal Society working group on

response to increased atmospheric CO2, Nature 407 (2000) Ocean acidification, personal communication. 364–367. [91] I. Zondervan, R.E. Zeebe, B. Rost, U. Riebesell, Decreasing Andy Ridgwell is an Assistant Professor marine biogenic calcification: a negative feedback on rising in the Department of Earth and Ocean

atmospheric pCO2, Glob. Biogeochem. Cycles 15 (2001) Sciences of the University of British 507–516. Columbia, and has the fancy title of [92] R.A. Armstrong, C. Lee, J.I. Hedges, S. Honjo, S.G. dCanada Research Chair in Global Proc- Wakeham, A new, mechanistic model for organic carbon ess ModellingT. Although in practice fluxes in the ocean: based on the quantitative association of spending most of his time tending to POC with ballast minerals, Deep-Sea Res., Part II 49 (2002) the every need of 5 cats, his research 219–236. addresses fundamental questions sur- [93] C. Klaas, D.E. Archer, Association of sinking organic matter rounding the past and future controls on

with various types of mineral ballast in the deep sea: atmospheric CO2 and the role of feed- implications for the rain ratio, Glob. Biogeochem. Cycles backs in the climate system. His weapon of choice in this endeavor is 16 (2002), doi:10.1029/2001GB001765. an Earth System Climate Model. [94] A.J. Ridgwell, An end to the drain ratioT reign?, Geochem. Geophys. Geosyst. 4 (2003), doi:10.1029/2003GC000512. Richard Zeebe is an Assistant Professor [95] S. Barker, J.A. Higgins, H. Elderfield, The future of the at the Department of at carbon cycle: review, calcification response, ballast and the University of Hawaii. His research

feedback on atmospheric CO2, Philos. Trans. R. Soc. A focuses on the global carbon cycle, 361 (2003) 1977. biogeochemistry and paleoceanography.

[96] C.L. Sabine, et al., The oceanic sink for anthropogenic CO2, His interest ranges from physico-chem- Science 305 (2004) 367–371. ical properties of molecules and the [97] G.R. Dickens, Rethinking the global carbon cycle with a biogeochemistry of foraminifera to large, dynamic and microbially mediated gas hydrate the scale of the global ocean. Together capacitor, EPSL 213 (2003) 169–183. with Dieter Wolf-Gladrow he has [98] S. Bains, R.D. Norris, R.M. Corfield, K.L. Faul, Termi- published a book in 2001 on the

nation of global warmth at the Palaeocene/Eocene boun- CO2 chemistry in seawater which has been referred to as dthe CO2 dary through productivity feedback, Nature 407 (2000) survival kitT. 171–174.