You are on page 1of 11

ARTICLE IN PRESS

Sedimentary Geology xx (2003) xxx xxx www.elsevier.com/locate/sedgeo

Transgressive surfaces of erosion as sequence boundary markers in cool-water shelf carbonates


Vincent Caron *, Campbell S. Nelson, Peter J.J. Kamp
Department of Earth Sciences, The University of Waikato, Private Bag 3105, Hamilton, New Zealand Received 19 August 2003; accepted 8 October 2003

Abstract Although sequence stratigraphy was first conceptualized in, and applied to, siliciclastic successions, its methodology and concepts are applicable to the carbonate regimen, provided that potential modifications to the existing models are appreciated that take into account the specificities of carbonate systems. We demonstrate that some properties inherent to isolated (i.e. continent-detached), shallow-water, temperate carbonate systems, including their typical low diagenetic potential due to the prevalance of calcitic mineralogies, the critical role exerted by storm waves and tidal currents on sediment dynamics (e.g. shedding of sediments at wave abrasion depth during highstand and lowstand conditions), and uneven sea-bed topographies cause both the genesis and preservation of subaerial unconformities to be uncommon, thereby rendering their use as sequence boundaries inappropriate. Instead, transgressive surfaces of erosion (TSEs) that formed as wave base shifts upwards during early rise of relative sea level, are readily identifiable in the field, extensive and, by their very nature, are preserved in the rock record, typically at the base of upward-deepening stacked facies. In Pliocene cool-water carbonate successions from New Zealand, the TSEs are characterized by starved sedimentation and burrowed network development, and are overlain by coarse bioclastic shoreface sediments and shell beds. In most cases, the surfaces top-truncate shoaling-upward deposits, but also deepening-upward stacking facies in places. Emersive events rarely correspond to actual physical surfaces, but diagenetic evidence preserved beneath TSEs indicates that occasionally subaerial surfaces had previously developed and were later modified into TSEs. Therefore, we propose that TSEs are the most adequate unconformities to define high-order depositional sequences (20 250 ka) in the studied successions because, by contrast with emersion surfaces, their formation and preservation are not dependent upon cool-water sedimentary facies and mineralogy, and are favoured by the dynamics of the setting (i.e. wave base shift and accommodation creation during relative sea-level rise). D 2003 Elsevier B.V. All rights reserved.
Keywords: Cool-water carbonates; Sequence stratigraphy; Transgressive surface of erosion; Sequence boundary; Depositional sequences

1. Introduction Complications can arise when trying to apply the sequence stratigraphic methodology, first conceptualized in siliciclastic successions, to carbonate settings because the controlling factors on the sequence stratigraphic development of the two systems may differ

* Corresponding author. 17 Rue dAmiens Domart S/La Luce 80110 France. E-mail addresses: simardvi@yahoo.com (V. Caron), c.nelson@waikato.ac.nz (C.S. Nelson). 0037-0738/$ - see front matter D 2003 Elsevier B.V. All rights reserved. doi:10.1016/j.sedgeo.2003.10.001

SEDGEO-03225; No of Pages 11

ARTICLE IN PRESS
2 V. Caron et al. / Sedimentary Geology xx (2003) xxxxxx

markedly both in their nature and effect (Sarg, 1988; Schlager, 1991,1993; Brachert et al., 2003). Accommodation and sediment fluxes are relatively independent in siliciclastic systems because sediments are produced outside the shelf, and enter the system depending on factors controlling the drainage from

continental sources. By contrast, in carbonate systems, rates of sediment production are more closely related to the formation and loss of accommodation on the shelf, especially to rates of sea-level rise and the nature of sea-floor topography. In addition, Schlager (1991,1999) pointed out that relative sea-level changes

Fig. 1. (A) Locality map of North Island, New Zealand, showing study region (arrowed) in forearc basin inboard of the convergent Pacific Australian plate boundary. (B) General distribution of the Plio Pleistocene limestones in the study area. In the western sector, limestones (in dark grey) are assumed to have formed on a narrow shelf attached to a landmass. By contrast, in the eastern sector, time-equivalent limestones (in black) are interpreted to have developed on upthrust-cored antiforms detached from a landmass (Caron, 2002). (C) Inferred depositional and sequence stratigraphic model for isolated carbonate systems as envisaged for the cool-water carbonate successions in eastern North Island. Carbonate factories were located upon submarine highs in relation to growth fold structures. Note that in cool-water carbonate systems where reefs are absent and early sea-floor cementation is poor, the depth to which the sea bed is stirred by fair-weather and storm waves (thick black arrows) is critical to sediment dispersal and sequence patterns, and determines the shape of the marine equilibrium profile. TSTtransgressive systems tract, HSThighstand systems tract, RSTregressive systems tract, FRSTforced regressive systems tract.

ARTICLE IN PRESS
V. Caron et al. / Sedimentary Geology xx (2003) xxxxxx 3

alone do not adequately explain stratal and facies stacking patterns in carbonate depositional systems. To accommodate these short-comings, he integrated in existing sequence stratigraphic models, factors such as nutrient supply, hydraulic energy, depth of light penetration, terrigenous input and platform type (see also Pomar, 2001) that specifically impact carbonate biotic associations, and in turn sediment production, and determine sediment dispersal. The resulting, more process-oriented, sequence stratigraphic framework, proves a powerful tool when applied to tropical carbonate successions, from which it originally derives, to assess and predict complex facies distribution and geometries (Schlager, 1992,1999; Moore, 2001). Here, we propose to extend this framework to nontropical carbonate systems since there exist some major differences between tropical and non-tropical carbonate-producing biota, both in their composition and sensitivity to the aforementioned environmental parameters (Nelson, 1988; James, 1997). It is demonstrated that types 1 and 2 sequence boundaries sensu Van Wagoner et al. (1988) cannot adequately be used to define most high-order depositional sequences in non-tropical carbonate successions (20 250 ka in our examples; Caron, 2002) because of inherent properties that generally do not permit the genesis or preservation of subaerial unconformities. Instead, we show that transgressive surfaces of erosion (TSEs) (Bhattarcharya, 1993) are the only satisfactory unconformities to hierarchically organize stratal and facies stacking patterns in cool-water carbonate successions. Shallow-water Pliocene cool-water limestones from eastern North Island, New Zealand, serve to illustrate our discussion (Fig. 1A and B), with an emphasis on some properties of cool-water carbonate systems.

2. Some considerations of cool-water carbonate systems The relatively reduced level of carbonate saturation in non-tropical shelf sea waters and the predominance of a low- to intermediate-Mg calcite mineralogy in the skeletal carbonate facies are responsible for a paucity of sea-floor cementation and an overall low chemical reactivity of non-tropical carbonate deposits (James, 1997; Caron and Nelson, 2003). This contrast in

mineral solubility with aragonite- and high-Mg calcite-rich tropical deposits is particularly critical to the record of meteoric influence during exposure at sequence boundaries. The occurrence of photozoan (light dependent; e.g. coralgal) biotic associations in tropical settings versus heterozoan (light independent; e.g. bryomol) assemblages in non-tropical shelf carbonates results in differing sensitivities to environmental factors (James, 1997). The efficiency of carbonate factories in coolwater environments dominated by filter/suspensionfeeding biota is limited by trophic resources and siliciclastic inputs. Consequently, a locus of deposition outside siliciclastic fairways and high-energy hydraulic conditions that prevent siliciclastic dilution and burial, and accelerate nutrient renewal, are underlying controls on promoting carbonate producers such as bryozoans and barnacles, provided that suitable substrates are also available (Henrich et al., 1995; Hayton et al., 1995; James, 1997). Because growth rates of temperate carbonate systems are typically low, from a few to 10 cm/ ky (Nelson, 1988; James, 1997), temperate carbonate systems are particularly sensitive to relative sealevel changes. In the case of isolated systems, a decrease in current velocities accompanying sealevel rise causes the suspended load to settle and swamp suspension filter feeders. Therefore, periods of limestone development are likely to occur when there is an equilibrium between the rates of subsidence and eustatic changes such that water depths are relatively uniform at production sites, thereby allowing carbonate biota to keep up with slow sealevel changes. In the absence of fauna capable of constructing wave-resistant structures, sediments in cool-water environments tend to be shed off the production site by high-energy waves and tidal currents. As a result, the space available to accommodate sediments is wedged between the depositional surface and the wave abrasion depth where tide and storm swells have the potential to winnow away fine particles and rework coarser material (Fig. 1C; James et al., 1994; Brachert et al., 2003). Therefore, inherited topography, tectonically and/or glacioeustatically induced sea-level changes, and climate are key factors governing changes in accommodation, and in turn the genesis and preservation of unconformities.

ARTICLE IN PRESS
4 V. Caron et al. / Sedimentary Geology xx (2003) xxxxxx

3. Pliocene examples of depositional sequences Examples presented here are drawn from Pliocene carbonate sediments from eastern North Island of New Zealand that formed in shallow temperate waters ( < 60 m) on isolated platforms developed as upthrust-cored submarine ridges that were swept by strong tidal flows and storm waves (Fig. 1C; Kamp et al., 1988). The successions are divided into between 4 and 12 packets (each 5 20 m thick), bounded by physical stratal surfaces and interpreted as depositional sequences that were generated

through the combined effects of tectonism and recurring high-frequency sea-level fluctuations (Caron, 2002; Caron and Nelson, 2003). Three major types of sequence motifs occur in the limestone units: (1) condensed sequences a few metres thick in up-dip locations on the depositional profile (Fig. 2A); (2) truncated sequences exhibiting a strong retrogradational (deepening-upwards) facies trend, and corresponding to sequences partly or totally amputated from, or lacking, a shoalingupward member (Fig. 2B); and (3) complete sequences having a lower retrogradational and

Fig. 2. Distribution of depositional sequence motifs upon and about the flanks of a submarine high (e.g. Fig. 1C), showing their component systems tracts, and the nature of physical surfaces bounding them. It is asserted that the development of these stratigraphic architectures is controlled by the inherited topography (i.e. troughs versus arches), the marine profile of equilibrium, a function of the depth of wave abrasion, and relative sea-level changes that generate erosive unconformities (i.e. TSE and RSE). Note that starvation facies occur typically at the base of the TST and towards the top of the RST, and at the TST-HST turnaround. Sequence motifs are based on observed lithologies (Caron, 2002).

ARTICLE IN PRESS
V. Caron et al. / Sedimentary Geology xx (2003) xxxxxx 5

an upper progradational (shallowing-upwards) facies trend, in down-dip positions or in topographic troughs (Fig. 2C). Complete sequences comprise the following facies and surfaces in ascending stratigraphic order (Fig. 3): (1) a transgressive systems tract (TST) starting with a basal ravinement surface, typically penetrated by Ophiomorpha, Skolithos and locally Thalassinoides burrows, overlain by a thin ( < 0.2 m) condensed bioclastic lag, with abundant detrital glauconite and mud clasts, followed by either shallow-water fossiliferous and pebbly rudstone, or coarse skeletal-rich cross-bedded shoreface-type facies. These are in turn overlain by finer-grained mixed carbonate-siliciclastic inner-shelf facies, and a condensed horizon, typically including planktic foraminifera; (2) above the TST follows a highstand systems tract (HST) consisting of aggradational (planar- and tabular-bedded) mixed carbonate-siliciclastic mid- to outer-shelf facies; and lastly (3) a regressive systems tract (RST; Naish and Kamp, 1997) that comprises progradational shallowwater cross-bedded skeletal shoreface-type facies. The transition from aggradational to progradational stratal geometries is used to distinguish highstand HST and RST. The RST is bounded above by a burrowed ravinement surface, which in rare cases bears discrete evidence of subaerial exposure. In a few examples, mid- to outer-shelf sediments of the HST or upper TST are overlain abruptly across an erosional surface by shallow-water shoreface sediments assigned to a forced regressive systems tract (FRST, b in Fig. 1C; Hunt and Tucker, 1995), in turn bounded above by a ravinement unconformity. Condensed sequences (Fig. 2A) comprise a thin TST ( < 1 m), which corresponds to a condensed shallow-water fossiliferous to pebbly rudstone. HST and RST are typically amalgamated in shallow-water cross-bedded grainy facies that include horizons with marine cements (Caron, 2002). On the basis of the depositional architecture displayed by the successions, the identified unconformities are assumed to coincide either with the transgressive surface of erosion (Bhattarcharya, 1993) at the base of upward-deepening deposits or regressive surfaces of erosion (Nummedal et al., 1993) at the base of forced regressive stacking facies. The absence or rarity of subaerial surfaces preserved in the stratigraphic record investigated here, and the

Fig. 3. Sequence-bounding unconformities: (A) contact (dashed line) interpreted as a TSE between medium-grained, well sorted, nearshore to beach deposits from the regressive systems tract of the previous sequence below and coarse-grained shoreface deposits at the base of the TST of the next sequence above. Hammer (arrowed) is about 30 cm long. (B) TSE (dashed line arrowed), with a relief up to 20 cm, between coarse bioclastic shoreface grainstone (base of overlying TST) and interbedded fine-grained sandstone and terrigenous mid- to outer-shelf packstone (upper TST of the underlying sequence). Person for scale at left. (C) Close up view of contact in (B). The surface (white arrows), bioturbated by Skolithos-type burrows (black arrow) and Scolicia (Sc), is interpreted as a TSE. Burrows are filled by mixed carbonatesiliciclastic skeletal hash from the overlying TST.

ARTICLE IN PRESS
6 V. Caron et al. / Sedimentary Geology xx (2003) xxxxxx

lack of lowstand deposits, leave little alternative but to use the transgressive surfaces of erosion to define depositional sequences (Figs. 1 and 2). This apparent inconsistency with traditional models is proposed to be inherent to cool-water carbonate systems (see later).

4. Key boundaries and horizons 4.1. Transgressive surface of erosion Transgressive surfaces of erosion develop when sea level rises and correspond to the process of erosional translation of the shoreface profile upwards and landwards (Nummedal and Swift, 1987), involving an intense reworking and winnowing of preexisting deposits by the action of wave and storm currents (Fig. 4C). TSEs have commonly been interpreted to form immediately after maximum regression of the shoreline (Bhattarcharya, 1993; Embry, 1995), and as such may also be responsible for the removal of regressive surfaces of erosion (RSEs) and surfaces of subaerial exposure (type 1 or type 2 sequence boundaries). TSEs are sharp or erosional with a relief up to 50 cm, and underlie deepening-upwards deposits (Fig. 3). This ravinement surface is characterized by a network of various ichnofossils, including Ophiomorpha, Skolithos and Thalassinoides that penetrate downward into older deposits. They are filled by coarse material of the overlying TST, and are indicative of periods of low or arrested sedimentation (Fig. 3C). The TSEs are frequent and easily identifiable in the field, and suggest that in many cases underlying deposits were top-truncated during transgressions (Fig. 3B). 4.2. Regressive surface of erosion

Swift, 1987; Walker and Wiseman, 1995). Erosion of the sea floor is due to storm wave action, considering that with lowered relative sea level storm wave base would be lowered as well. Thus, forced regressive deposits are underlain by a ravinement surface equivalent to a regressive surface of marine erosion (Figs. 1C and 4B) produced through wave planing during sea-level fall (Posamentier et al., 1992), or a correlative conformity down-dip of the area where the erosional surface forms. The regressive deposits are detached from the previous highstand shoreline and the RSE may be connected upward or landward to a surface of subaerial exposure. It would then coincide with a type 2 sequence boundary (Van Wagoner et al., 1988). It is expected that during late sea-level fall, sediments of the previous sequence deposited during sea-level highstand (HST) or during an earlier phase of sealevel fall (RST) would be partly or totally removed through wave planing, potentially later overprinted by a TSE (Figs. 1C and 4C). RSEs are distinct from wave-cut erosion surfaces formed during normal regression in up-dip position of the depositional profile at the depth of wave abrasion (James et al., 1994; Boreen and James, 1995; Brachert et al., 2003). In conditions of slow rates of sea-level rise and high rates of carbonate production, space created for sediment accumulation may not be sufficient to accommodate sediments rapidly aggrading to the depth of wave abrasion. As a result, the excess of sediment is by-passed basinward by currents and a wave-cut erosion surface is generated (Fig. 4A). This process, referred to by Schlager (1992, 1993) as highstand shedding of carbonates, is likely to be particularly effective in cool-water carbonate settings where little sea-floor cementation and the absence of wave-resistant reefs prevent stabilization of sediments. 4.3. Subaerially exposed surface

The abrupt occurrence of sharp-based, anomalously coarser and more proximal shallowing-upwards deposits in a more distal marine setting are the manifestation of a basinward and downward shift of the coastal environment, otherwise commonly interpreted as a forced regression, which occurs when the rate of relative sea-level fall exceeds the rate of subsidence of the host basin (Nummedal and

The loss of accommodation space, due to decreasing bathymetry associated with high rates of carbonate production and falling sea level, causes the eustatic sea level to reach the elevation of the sea floor, and a subaerial unconformity forms (Van Wagoner et al., 1988). In isolated carbonate systems, subaerial exposure is responsible for the development of meteoric

ARTICLE IN PRESS
V. Caron et al. / Sedimentary Geology xx (2003) xxxxxx 7

Fig. 4. Schematic depositional and time-stratigraphic diagrams illustrating the effect of wave abrasion and successive erosional processes that accompany the development of discontinuity surfaces, in relation to relative sea-level changes, on sediment dispersal and preservation of previously developed systems tracts and surfaces. In (A), the model emphasizes that in up-dip positions of the depositional profile, sediments produced close to the depth of wave abrasion cannot be accommodated and are shed downslope (black arrows indicate the winnowing effect on unarmoured sediments). In this case, the sedimentary record of time is null on the antiform tops. In (B), a relative sea-level fall causes the lowering of abrasion wave base. As a result, sediments are removed in accordance with the formation of a RSE and accumulate in the form of a FRST. In (C), the model shows that an actual physical TSE, developed during the ensuing transgression, has recorded many metres of erosion, and overprinted previous surfaces, including RSEs and potentially subaerial exposure surfaces. Note the diachronous nature of the unconformities: RSEs are getting younger basinward, whereas TSEs are upward-younging.

lenses floating on marine water beneath the antiform caps. The extent of such lenses is determined by the porosity and permeability of the surficial sediment and rock, and the hydraulic potential that is controlled

primarily by relief on the watertable (Vacher, 1988). Considering the expected low relief on temperate carbonate sand islands, and both the low chemical reactivity of the calcitic mineralogies and the low rate

ARTICLE IN PRESS
8 V. Caron et al. / Sedimentary Geology xx (2003) xxxxxx

of fluid interaction upon carbonate grains, the duration of exposure to meteoric water systems becomes critical to the development of mouldic porosity and calcite cementation in cool-water environments. Subaerial exposure surfaces are rarely evident in the Pliocene successions investigated. In any case, never does a subaerial unconformity correspond to an actual physical surface. Thus, the recognition of subaerial exposure is always indirect, yielded by diagenetic evidence recorded beneath sharp to erosional surfaces interpreted as TSEs (Caron and Nelson, 2003), which are assumed to have removed any direct physical evidence of the antecedent subaerial unconformity (Figs. 2B and 4C). 4.4. Condensed horizon (Cdh) at TST-HST turnaround Within the studied depositional sequences, a thin ( < 1 m), locally sharp-based, mixed carbonate-siliciclastic horizon consisting of fine-grained reworked shell fragments and mud clasts, whose diagenesis indicates prolonged residence on a starved sea floor (Caron and Nelson, 2003), lies at the TST-HST turnaround (i.e. at the transition between retrogradational and aggradational-progradational stacking patterns). Starvation requires that clastic input be minimal relative to accommodation creation, either because sediments were trapped farther landward (i.e. the locus of terrigenous deposition is displaced in the nearshore environment) or the carbonate factories were partly drowned resulting in low carbonate production (Kidwell, 1991a). The condensed horizon (Cdh) rarely corresponds to wellindividualized shell beds bounded above by a downlap surface as defined by Kidwell (1991a,b), and documented for other Plio Pleistocene New Zealand examples by Abbott and Carter (1994) and Naish and Kamp (1997). It has indistinguishable boundaries, and the transition between TST and HST is assumed to lie somewhere within the horizon at some indeterminate level (Fig. 2). Recognition of the Cdh is based on sedimentologic and diagenetic evidence. Its traceability is generally poor, except when directly associated with easily correlatable shell beds located between mid- to outer-shelf siltstones and mudstones found above transgressive lags.

5. Discussion: key surfaces and sequence boundary Current difficulties in the sequence stratigraphic analysis of carbonate successions lie in uncertainties regarding the hierarchical significance of physical unconformities in outcrops. Posamentier et al. (1992) and others have argued that sequence boundaries should be placed at the base of the deposits that formed during sea-level falls (LST), and correspond to regressive surfaces of erosion and to the subsequent subaerial unconformities generated landward. However, the lack or poor traceability of subaerial unconformities in the studied Pliocene successions, only found in a few examples at the top of RSTs, renders their subdivision into depositional sequences sensu stricto difficult. As previously stated, the efficiency of carbonate factories, sediment dispersal, and subsequent sequence patterns are closely dependent upon environmental parameters that may also influence the genesis and preservation of unconformities. For example, the geometry of isolated carbonate depositional systems generates uneven sea-floor topographies where evidence of any relative sea-level fluctuation may only be recorded in quite restricted portions of the carbonate realm, especially in areas with high rates of sediment supply and tectonic subsidence. Elsewhere, such evidence may be more subtle or missing. Further, in cool-water carbonate systems, considering the typical low chemical reactivity of calcitic sediments, the genesis of emersion surfaces is a function of the amplitude of sea-level fall that controls the size of emerged lands, and its duration. Previous workers have suggested that the surface underlying transgressive deposits is the most appropriate surface to choose for bounding a sequence. Arnott (1995), Hunt and Tucker (1995) and Naish and Kamp (1997) placed sequence boundaries above regressive deposits (either forced or normal regressive systems tracts) and beneath transgressive deposits because the transgressive surface of erosion is typically extensive, readily identifiable in sections (cores or outcrops), and, by its very nature, usually preserved in the stratigraphic record (Fig. 4). Walker (1995) emphasized that falling sea levels may not always be preserved in the geological record and that instead many sequences might be bounded by unconformities developed during rising stages. Schlager (1999), with reference to types 1 and 2

ARTICLE IN PRESS
V. Caron et al. / Sedimentary Geology xx (2003) xxxxxx 9

sequence boundaries defined by Vail et al. (1977, 1984), proposed to acknowledge the flooding surface between a HST and the overlying TST, without evidence of sea-level fall or lowstand deposits in between, as a third boundary considering its particular usefulness in carbonates. Nummedal and Swift (1987) argued that because of its genesis as a surface of sediment transfer during retrogradational shoreface retreat, the TSE is diachronous, getting younger landward (Fig. 4C). Although the TSE may form a sharp, widespread lithologic boundary in the stratigraphic record, Demarest and Kraft (1987) showed that this surface should not be regarded as a hiatus between depositional sequences when it does not coincide with the RSE. However, the sharp and erosional surfaces that we recognize in the field overlain by deepening-upwards deposits indicate that, if RSE had previously developed (at the top of the RST according to Naish and Kamp, 1997), they were not preserved but instead were modified into TSEs. If these erosion surfaces were misinterpreted as TSEs and instead represent sequence boundaries (Van Wagoner et al., 1988; Van Wagoner, 1996), it might imply that the observed erosion associated with coarse-grained lags reflects valley incision and consequent by-pass of skeletal sands into the basin. In the successions investigated, there is no evidence for significant volumes of sediments by-passed basinward along the erosion surfaces (i.e. there are no lowstand deposits). Therefore, the erosion/sharp surfaces interpreted to have formed by transgressive ravinement are regarded as the most adequate to subdivide packages of sediments into depositional sequences, and as such are inferred to be potentially superposed on, or to have eroded through, previous RSEs hierarchically not necessarily equivalent to SB sensu Van Wagoner et al. (1988). The transgressive surfaces of erosion do not necessarily truncate regressive deposits. While commonly there is an abrupt grain-size change across the TSE defined by a coarse-grained transgressive lag, TSEs do not always mark the boundary between underlying shallowing-upwards facies and overlying deepening-upwards facies (see truncated sequencetype in Fig. 2B). Consequently the enclosed sequences are not strictly similar to the transgressive-regressive sequence of Embry (1995), as he also used TSEs as the main bounding discontinuity to define deposi-

tional units. In the cases where the TSE has eroded through a RST with preserved evidence of subaerial exposure, it then coincides with a sequence boundary and consists of a composite unconformity surface.

6. Conclusions Changes in accommodation, sediment supply and hydrodynamic regime of differing amplitude and duration are critical and interdependent factors controlling the style of depositional facies stacking, the genesis of key surfaces, diagenesis, stratal geometries and at a higher hierarchical level the piling up of sequences (Van Wagoner et al., 1988; Vail et al., 1991; Tucker, 1993; Pomar, 2001). In carbonate settings, sediment dispersal, facies stacking patterns, sequence architectures and the genesis and preservation of unconformities are also dependent upon carbonate production, early cementation, biotic association distributions and their dominant original mineralogies. In these regards, cool-water carbonate settings differ markedly from their tropical counterparts: rates of carbonate production are usually lower, biological binding is limited, early cementation is rare and the dominant stable calcitic mineralogy (i.e. low- and intermediate-Mg calcite) causes the diagenetic potential of sediments to be low. As a result: (1) cool-water carbonate factories tend to be swamped by rapid sealevel rise; (2) sediment dynamics and dispersal are controlled by wave abrasion processes; and (3) the genesis of physical subaerial unconformities, limited by the poor reactivity of sediments, depends upon the duration of exposure to meteoric processes, and climate (i.e. humid vs. arid). With this in mind, and on the basis of field observations, we propose that because environmental factors do not favour the preservation of subaerial unconformities and regressive surfaces of erosion in isolated cool-water carbonate settings, transgressive surfaces of erosion that develop during upward shift of wave base in the early stages of relative sea-level rise are the most adequate unconformities to bound depositional sequences. TSEs are readily identifiable in the field, typically extensive and, by their very nature, are preserved in the rock record at the base of upward-deepening stacked facies. They are characterized by starved sedimentation and burrowed network development,

ARTICLE IN PRESS
10 V. Caron et al. / Sedimentary Geology xx (2003) xxxxxx sequence stratigraphy and diagenetic pathways of cool-water shelf carbonate facies. Unpublished PhD thesis The University of Waikato, Hamilton, 445 p. Caron, V., Nelson, C.S., 2003. Developing concepts of highresolution diagenetic stratigraphy for cool-water limestones: application to Pliocene Te Aute limestones, New Zealand, and their sequence stratigraphy. Carbonates and Evaporites 18, 63 85. Demarest, J.M., Kraft, J.C., 1987. Stratigraphic record of Quaternary sea levels: implications for more recent strata. In: Nummedal, D. (Ed.), Sea-level Fluctuation and Coastal Evolution. SEPM Special Publication, vol. 41. SEPM (Society for Sedimentary Geology), Tulsa, Oklahoma, pp. 223 240. Embry, A.F., 1995. Sequence boundaries and sequence hierarchies: problems and proposals. In: Steel, R.J., Felt, V.L., Johannesen, E.P., Mathieu, C. (Eds.), Sequence Stratigraphy: Advances and Applications for Exploration and Production in Northwest Europe. Stavanger, Elsevier, Amsterdam, pp. 1 11. Hayton, S., Nelson, C.S., Hood, S.D., 1995. A skeletal assemblage classification system for non-tropical carbonate deposits based on New Zealand Cenozoic limestones. Sedimentary Geology 100, 123 141. Henrich, R., Freiwald, A., Betzler, C., Bader, B., Schafer, P., Samt leben, C., Brachert, T.C., Wehrmann, A., Zankl, H., Kuhlmann, D.H.H., 1995. Controls on modern carbonate sedimentation on warm-temperate to arctic coasts, shelves and seamounts in the northern hemisphere: implications for fossil counterparts. Facies 32, 71 108. Hunt, D., Tucker, M.E., 1995. Stranded parasequences and the forced regressive wedge systems tract: deposition during base level fallreply. Sedimentary Geology 95, 147 160. James, N.P., 1997. The cool-water carbonate depositional realm. In: James, N.P., Clarke, J.A.D. (Eds.), Cool-Water Carbonates. SEPM Special Publication, vol. 56. SEPM (Society for Sedimentary Geology), Tulsa, Oklahoma, pp. 1 22. James, N.P., Boreen, T.D., Bone, Y., Feary, D.A., 1994. Holocene carbonate sedimentation on the west Eucla Shelf, great Australian Bight: a shaved shelf. Sedimentary Geology 90, 161 177. Kamp, P.J.J., Harmsen, F.J., Nelson, C.S., Boyle, S.F., 1988. Barnacle-dominated limestone with giant cross-beds in a non-tropical, tide swept, Pliocene forearc seaway Hawkes Bay, New Zealand. Sedimentary Geology 60, 173 195. Kidwell, S.M., 1991a. Condensed deposits in siliciclastic sequences: expected and observed features. In: Einsele, G., Ricken, W., Seilacher, A. (Eds.), Cycles and Events in Stratigraphy. Springer, Berlin, pp. 682 695. Kidwell, S.M., 1991b. Taphonomic feedback (live/dead interactions) in the genesis of bioclastic beds: keys to reconstructing sedimentary dynamics. In: Einsele, G., Ricken, W., Seilacher, A. (Eds.), Cycles and Events in Stratigraphy. Springer, Berlin, pp. 268 282. Moore, C.H., 2001. Carbonate reservoirsporosity evolution and diagenesis in a sequence stratigraphic framework. Developments in Sedimentology, vol. 55. Elsevier, Amsterdam. 444 pp. Naish, T.R., Kamp, P.J.J., 1997. Sequence stratigraphy of sixthorder (41 k.y.) Pliocene Pleistocene cyclothems, Wanganui ba-

and are overlain by coarse bioclastic shoreface sediments and shell beds. As a concluding remark, we re-emphasize as did Schlager (1992,1993) that the understanding of the role exerted by environmental parameters on carbonate production and sediment dispersal is the key to the future development of predictive depositional models in carbonate settings and is especially applicable for successful sequence stratigraphic interpretations of cool-water carbonate successions. Acknowledgements We thank especially Alan Beu (Lower Hutt), Arne Pallentin (Waikato) and Kyle Bland (Waikato) for discussions in the field about the sequence stratigraphy of the eastern North Island Pliocene limestones. Funding support from the Marsden Fund contract UOW801 (Royal Society of New Zealand) and the Foundation for Research Science and Technology contract UOW608 to the University of Waikato is gratefully acknowledged. References
Abbott, S.T., Carter, R.M., 1994. The sequence architecture of MidPleistocene (c. 1.1 0.4 Ma) cyclothems from New-Zealand: facies development during a period of orbital control on sealevel cyclicity. In: de Boer, P.L., Smith, D.G. (Eds.), Orbital Forcing and Cyclic Sequences. International Association of Sedimentologists Special Publication, vol. 19. Blackwell Publishing, pp. 367 394 Arnott, R.W.C., 1995. The parasequence definitionare transgressive deposits inadequately addressed? Journal of Sedimentary Research B65, 1 6. Bhattarcharya, J.P., 1993. The expression and interpretation of marine flooding surfaces and erosional surfaces in core; examples from the upper Cretaceous Dunvegan Formation, Alberta Foreland basin, Canada. In: Posamentier, H.W., Summerhayes, C.P., Haq, B.U., Allen, G.P. (Eds.), Sequence Stratigraphy and Facies Associations. International Association of Sedimentologists Special Publication, vol. 18, Blackwell Publishing, pp. 125 160. Boreen, T.D., James, N.P., 1995. Stratigraphic sedimentology of tertiary cool-water limestones SE Australia. Journal of Sedimentary Research B65, 142 159. Brachert, T.C., Forst, M.H., Pais, J.J., Legoinha, P., Reijmer, J.J.G., 2003. Lowstand carbonates, highstand sandstones? Sedimentary Geology 155, 1 12. Caron, V., 2002. Petrogenesis of Pliocene limestones in southern Hawkes Bay, New Zealand: a contribution to unravelling the

ARTICLE IN PRESS
V. Caron et al. / Sedimentary Geology xx (2003) xxxxxx sin, New Zealand: a case for the regressive systems tract. Geological Society of America Bulletin 109, 978 999. Nelson, C.S., 1988. An introductory perspective on non-tropical shelf carbonates. Sedimentary Geology 60, 3 12. Nummedal, D., Swift, D.J.P., 1987. Transgressive stratigraphy at sequence-bounding unconformities: some principles derived from Holocene and Cretaceous examples. In: Nummedal, D., Pilkey, O.H., Howard, J.D. (Eds.), Sea-level Fluctuation and Coastal Evolution. SEPM Special Publication, vol. 42, pp. 358 370. Nummedal, D., Riley, G.W., Templet, P.L., 1993. High-resolution sequence architecture: a chronostratigraphic model based on equilibrium profile studies. In: Posamentier, H., Summerhayes, C.P., Haq, B.U., Allen, C.P. (Eds.), Sequence Stratigraphy and Facies Association. SEPM Special Publication, vol. 18, pp. 55 68. Pomar, L., 2001. Types of carbonate platforms: a genetic approach. Basin Research 13, 313 334. Posamentier, H.W., Allen, G.P., James, D.P., Tesson, M., 1992. Forced regressions in a sequence stratigraphic framework: concepts, examples, and exploration significance. American Association of Petroleum Geologists Bulletin 76, 1687 1709. Sarg, J.F., 1988. Carbonate sequence stratigraphy. In: Wilgus, C.K., Posamentier, H., Ross, C.A., Kendall, C.G.S.C. (Eds.), SeaLevel Changes: An Integrated Approach. SEPM Special Publication, vol. 42. SEPM (Society for Sedimentary Geology), Tulsa, Oklahoma, pp. 155 181. Schlager, W., 1991. Depositional bias and environmental change important factors in sequence stratigraphy. Sedimentary Geology 70, 109 130. Schlager, W., 1992. Sedimentology and sequence stratigraphy of reefs and carbonate platforms. American Association of Petroleum Geologists Continuing Education Course Note Series 34 Tulsa, OK. 71 pp. Schlager, W., 1993. Accommodation and supplya dual control on stratigraphic sequences. Sedimentary Geology 86, 111 136. Schlager, W., 1999. Type 3 sequence boundary. In: Harris, P.M., Saller, A.H., Simo, J.A. (Eds.), Advances in Carbonate Sequence Stratigraphy: Application to Reservoirs, Outcrops and Models. SEPM Special Publication, vol. 63. SEPM (Society for Sedimentary Geology), Tulsa, Oklahoma, pp. 35 45. Tucker, M.E., 1993. Carbonate diagenesis and sequence stratigra11 phy. In: Wright, V.P. (Ed.), Sedimentology Review, vol. 1. Blackwell, Oxford, pp. 51 72. Vacher, H.L., 1988. Dupuit Ghyben Herzberg analysis of stripisland lenses. Geological Society of America Bulletin 100, 580 591. Vail, P.R., Audemard, F., Bowman, S.A., Eisner, P.N., Perez-Cruz, C., 1991. The stratigraphic signatures of tectonics, eustasy and sedimentologyan overview. In: Einsele, G., Ricken, W., Seilacher, A. (Eds.), Cycles and Events in Stratigraphy. Springer, Berlin, pp. 617 659. Vail, P.R., Mitchum Jr., R.M., Todd, R.G., Widmier, J.M., Thompson III, S., Sangree, J.B., Hatlelid, W.G. 1977. Seismic stratigraphy and global changes of sea level. In: Payton, C.E. (Ed.), Seismic StratigraphyApplication to Hydrocarbon Exploration. American Association of Petroleum Geologists Memoir, vol. 26, pp. 49 212. Vail, P.R., Hardenbol, J., Todd, R.G., 1984. Jurassic unconformities, chronostratigraphy, and sea-level changes from seismic stratigraphy and biostratigraphy. In: Schlee, J.S. (Ed.), Interregional Unconformities and Hydrocarbon Accumulation. American Association of Petroleum Geologists Memoir, vol. 36, pp. 129 144. Van Wagoner, J.C., 1996. Overview of sequence stratigraphy of foreland basin deposits: terminology, summary of papers, and glossary of sequence stratigraphy. American Association of Petroleum Geologists Memoir 64, ix xxi. Van Wagoner, J.C., Posamentier, H.W., Mitchum, R.M.J., Vail, P.R., Sarg, J.F., Loutit, T.S., Hardenbol, J., 1988. An overview of the fundamentals of sequence stratigraphy and key definitions. In: Wilgus, C.K., Hastings, B.S., Kendall, C.G.St.C., Posamentier, H.W., Ross, C.A., Van Wagoner, J.C. (Eds.), Sea-Level Changes: An Integrated Approach. SEPM Special Publication, vol. 42. SEPM (Society for Sedimentary Geology), Tulsa, Oklahoma, pp. 39 45. Walker, R.G., 1995. Sedimentary and tectonic origin of a transgressive surface of erosion: Viking Formation Alberta, Canada. Jounal of Sedimentary Research B65, 209 221. Walker, R.G., Wiseman, T.R., 1995. Lowstand shorefaces, transgressive incised shorefaces, and forced regressions: examples from the Viking Formation Joarcam area, Alberta. Journal of Sedimentary Research B65, 132 141.

You might also like