Você está na página 1de 11

Quaternary Research 54, 163173 (2000)

doi:10.1006/qres.2000.2164, available online at http://www.idealibrary.com on

Whole-Rock Aminostratigraphy and Quaternary Sea-Level History


of the Bahamas
Paul J. Hearty
School of Earth Sciences, James Cook University, Townsville, Queensland 4811, Australia

and
Darrell S. Kaufman
Departments of Geology and Environmental Sciences, Northern Arizona University, Flagstaff, Arizona 86011-4099
Received November 9, 1999

chronology, the sequence of depositional events can be defined


on the basis of stratigraphic order and characterized by unique
physical properties of each of the limestonesoil units.
The Bahama Islands are crucial to sea-level studies because
they are tectonically stable (Carew and Mylroie, 1995) and
thus provide direct evidence of global ice-volume changes.
Furthermore, countless outcrops are accessible on the nearly
700 Bahama Islands. Previously, however, direct dating of
these outcrops has been hampered by the lack of suitable
methods and materials. Only a small percentage of the Quaternary marine deposits in the Bahamas contains corals for
uranium-series dating, and the majority of the coral-rich deposits date to marine isotope substage (MIS) 5e. Radiocarbon
dating is only applicable to deposits 40,000 yr old, leaving a
dating gap for non-Holocene or noncoraliferous high sealevel deposits.
The utility of amino acid racemization (AAR) geochronology has been demonstrated in numerous studies (Rutter and
Blackwell, 1995). The whole-rock method was introduced in
Bermuda (Hearty et al., 1992) and has since been successfully
applied to Quaternary shorelines of the Bahamas (Kindler and
Hearty, 1997) and Hawaii (Hearty et al., 2000). Mitterer (1968)
showed that coated grains (i.e., ooids and peloids) and aragonite mud contain concentrations of amino acids similar to those
in marine and terrestrial mollusks, the more traditional sample
materials. Whether skeletal or oolitic grains, organic residues
in whole-rock samples show similar kinetic behavior, and yield
consistent amino acid ratios both stratigraphically and geographically.
The objectives of this investigation are: (1) to summarize,
independent of the aminostratigraphy, a physical stratigraphy
of the Bahama Islands; (2) to use whole-rock aminostratigraphy for confirmation and correlation of several rock units
throughout the Bahama Islands; and (3) to highlight information regarding sea-level history during the past several glacial
interglacial cycles.

The surficial geology of the tectonically stable Bahamian archipelago preserves one of the most complete records of middle to late
Quaternary sea-level-highstand cycles in the world. However, with
the exception of deposits from marine isotope substage (MIS) 5e,
fossil corals for radiometric dating of this rich stratigraphic sequence are rare. This study utilizes the previously published,
independent lithostratigraphic framework as a testing ground for
amino acid racemization in whole-rock limestone samples. At least
six limestonesoil couplets provide a relative age sequence of
events that encompass as many interglacial glacial cycles.
D-Alloisoleucine/L-isoleucine data fall into six clusters, or aminozones. On the basis of independent dating and the inferred
correlation with global MIS, the ages of several aminozones are
known, while the ages of others are calculated from calibrated
amino acid geochronology. This study demonstrates the utility of
the whole-rock aminostratigraphy method for dating and correlating widespread emergent marine deposits, constitutes the first
regional geochronological framework for the Bahamas, and highlights major sea-level events over the past half million
years. 2000 University of Washington.

INTRODUCTION

The mobile sedimentary blanket on the flat-topped Bahama


Platform (Fig. 1) is a sensitive monitor of sea-level fluctuations. Waves and wind, driven by powerful storms, remold the
sediments into beach, dune, and washover ridges, which upon
subaerial exposure are frozen in place by rapid cementation.
Carbonate sedimentation occurs during broad interglacial highstands while the platform is partially or wholly submerged, and
karst and soil processes alter these deposits during lowstands.
Major cycles, as well as minor fluctuations of sea level, are
thus recorded in the stratigraphic record. Independent of any
Supplementary data for this article may be found on the journal home page
(http://www.academicpress.com/qr).
163

0033-5894/00 $35.00
Copyright 2000 by the University of Washington.
All rights of reproduction in any form reserved.

164

HEARTY AND KAUFMAN

FIG. 1. Map of the Bahama Islands. A 200-m bathymetric contour outlines shelf areas of the Bahama Banks, Florida, and Cuba. Sample sites from the
Appendix are identified by island code (e.g., G, Grand Bahama; A, Abaco) and number.

APPROACH AND METHODS

The Quaternary stratigraphy of the Bahamas consists of a


succession of interglacial limestone facies separated by glacialage terra rossa paleosols. A sequence of stacked limestone
red-soils unitsthe younger superimposed upon the olderis
the most fundamental and independent measure of relative age.
This sequence exists in its entirety, as in north Eleuthera
(Hearty, 1998), and in part (Hearty and Kindler, 1993a,b, 1997;
Kindler and Hearty, 1997) at several islands throughout the
Bahamas.
We interpret the carbonate sedimentary facies (dune, beach,
and subtidal) as equivalent to interglacial highstand events and
terra rossa (i.e., reddish silty or clayey soils, generally 7.5 YR
5/3 or redder) paleosols as a record of intervals of lower-thanpresent sea level during midglacial to glacial intervals. In many
places, several of these limestonesoil couplets are stacked in
a vertical succession clearly observable in outcrop (Fig. 2).
Under closer examination, each of these couplets displays a
unique set of physical characteristics that distinguish them
from older and younger couplets. These include an assemblage
of sedimentary structures and facies, alternating limestone
composition, and evolving soil color, texture, and degree of
karstification and diagenesis. Given this observable succession
of interglacial glacial couplets, we infer, counting backward,

that the stratigraphically youngest was deposited during the


present interglaciation (MIS 1), that the next oldest is the last
interglaciation (MIS 5), and on back several interglaciations to
the oldest known couplet which must equal or exceed the age
of MIS 13. In addition to the physical stratigraphy, independent
radiometric ( 14C and U-series) ages from other studies reinforce our inferred correlation with isotope stages, or constrain
the ages of several of these couplets (i.e., MIS 1, 5e, 9, 13,
and 13). We prefer to identify units in association with
marine isotope stages, rather than to add to the already confusing nomenclature.
Over the past decade, more than 500 sites in Grand Bahama,
Abaco, Andros, New Providence, Eleuthera, Cat, Exumas, San
Salvador, Long, Acklins, Mayaguana, Caicos, and Great Inagua islands (Fig. 1) were investigated. Previous studies have
examined the stratigraphy and carbonate petrography of many
of these sites (e.g., Kindler and Hearty, 1996), from which it
was determined that sea-level modulation of sediment formation has similarly and simultaneously affected all the islands,
resulting in a collection of either oolitic or skeletal rocks. Local
processes that may alter sediment type appear to be of secondary importance. From within this regional stratigraphic framework, 275 representative whole-rock samples (Appendix,
available as supplementary data) were analyzed for AAR.

BAHAMAS STRATIGRAPHY AND SEA-LEVEL HISTORY

165

FIG. 2. Quaternary stratigraphy of north Eleuthera (see Hearty (1998) for a more detailed analysis). (A) Lithostratigraphy showing mean A/I ratios and
limestone composition (oolitic, peloidal, agal, and skeletal). Site locations from the Appendix are numbered in boxes. (B) Schematic synthesis identifying
limestonesoil couplets (IVI). Inferred correlation with marine isotope stages is based on the physical stratigraphy and succession of these couplets.

The underlying theory and various applications of the AAR


method are summarized in Rutter and Blackwell (1995). The
ratio of D/L (D-alloisoleucine/L-isoleucine, or A/I) amino acids
measures the extent of racemization, or epimerization in the
case of isoleucine. In living organisms, the A/I ratio is initially
zero (0.01 with laboratory preparation) and increases to an
equilibrium A/I ratio of about 1.3. Because the whole-rock
method analyzes aggregates of skeletal or coated grains that
form over time (i.e., having an inherited age), entirely modern material is not expected. In addition to independent molecular factors, the rate of the isoleucine epimerization/
racemization reaction depends on the ambient temperature
within the geologic deposit; thus, the epimerization rate at
more southerly sites is expected to be somewhat greater than at
higher latitudes. Present mean annual temperature (MAT)
serves as a less-than-ideal proxy for integrated thermal history.

Sample Preparation and Analysis


Whole-rock sample preparation procedures are outlined in
Hearty (1998). Samples were gently disaggregated and sieved
to obtain the 250- to 850-m fraction. This size range excludes
fine intergranular cements as well as large grains that could
significantly influence the A/I ratio. Cements have the potential
to yield significantly lower A/I ratios (Appendix, Abaco samples 1104A,D). Microscopic examination confirms the separation of grains from cements.
Sample analyses performed at Utah State University and
Northern Arizona University followed standard procedures
prescribed in Miller and Brigham-Grette (1989). Approximately 100 mg of each sample was first leached in dilute
hydrochloric acid to remove 30% of the sample weight and
reduce the possibility of contamination by removing any re-

166

HEARTY AND KAUFMAN

maining cements or other organic residues on grain surfaces.


Approximate 30-mg samples were dissolved in 7 M HCl containing 6.25 M norleucine (a nonprotein amino acid used as
an internal standard). Samples were flushed with N 2, sealed in
sterile vials, hydrolyzed at 110C for 22 h, and then evaporated
under N 2 in a heat block or under vacuum. After rehydration,
samples were injected onto an ion-exchange HPLC that employs postcolumn derivitization in O-phthaldialdehyde and fluorescence detection.
Each sample solution was analyzed three to five times and
the results were averaged. Analytical error from multiple analyses on the same sample solution is reported only in the
Appendix. The analytical precision of peak-height A/I ratios
was typically 3%. To monitor analytical drift and facilitate
comparison with other laboratories, the Amino Acid Geochronology Laboratory routinely measures the Interlaboratory
Comparative Standards of Wehmiller (1984) (Table 1, footnote).
In order to assess the effective age range and limitations of
the AAR method, numerous less-than-ideal samples were analyzed. Ultimately, based primarily on low amino acid concentrations, the ratios from 41 of the 275 samples (15%) were
excluded (Appendix). Two-thirds (27) of the spurious results
can be attributed to excessively low concentrations of amino
acids. With increasing age and recrystallization from aragonite
to calcite, amino acids are lost, apparently by leaching and
decomposition. Six samples were excluded on the basis of
proximity to soil horizons. The remaining eight rejected samples clearly violated stratigraphic order for unknown reasons.
These could be the result of extensive sediment reworking,
unknown chemical processes in the deposit, or minor laboratory inconsistencies.
BAHAMAS AMINOSTRATIGRAPHY

Couplet I Limestone, MIS 13?


The oldest surficial deposits now recognized in the Bahamas
lie at the base of the stratigraphic section at Goulding Cay,
Eleuthera (Fig. 2; EGC, sections 7, 2A, and 2B). These rocks
are typically recrystallized skeletal eolianites capped by multiple, intercalated, and complex deep-red paleosols. Two units
separated by a minor paleosol are identifiable within this massive eolianite (Fig. 2, base of section 2A). No intertidal facies
were observed. Most samples from this unit yielded low concentrations of amino acids, but one sample from Goulding Cay
has an A/I ratio of 0.79 (n 1). Another ratio of 0.74 was
derived from one of several wave-deposited boulders (Hearty,
1998) of skeletal composition. Although these two samples are
not sufficient to establish a firm relative age, the basal stratigraphic position of this unit and overlying ages (by thermal
ionization mass spectrometry, TIMS) suggest that the deposit
is MIS 13.

Couplet II Limestone, MIS 9/11


Middle Pleistocene interglacial deposits of MIS 9 and/or 11
are characterized by their partially calcitized, oolitic to oolitic/
peloidal composition (Kindler and Hearty, 1996), and their
extensive and complex coastal ridges containing subtidal,
beach, and eolian facies. The capping deep-red clayey paleosols fill karst pits up to 2 m deep. Near Goulding Cay, north
Eleuthera (Fig. 2, and Hearty, 1998) an assemblage of four
soil-bounded and/or unconformable facies units occurs in
stratigraphic succession within Couplet II. This sequence rests
directly beneath four younger glacialinterglacial couplets
(IIIVI) correlated with MIS 7, 5e, 5a, and 1 (Hearty, 1998;
this study), respectively.
Thirty-six AAR samples of this oolite-rich unit collected
across the Bahamas yielded consistent A/I ratios averaging
0.68 0.03 (Table 1). On New Providence Island (Hearty and
Kindler, 1997), deposits of MIS 9 and 11 are distinguished by
distinct high ridges that appear to separate deposits of the two
interglaciations, with the younger ridge yielding a 300,000-yr
conventional uranium-series age estimate on oolite (Muhs et
al., 1990). In Eleuthera (Fig. 2, section 2C), TIMS uranium
ages from a slightly altered oolite sample from the upper part
of this sequence returned an age of 550,000 yr (Hearty et al.,
1999a). Although a relatively new application, Muhs et al.
(1994) have demonstrated measured success with TIMS dating
of oolitic limestone.
Couplet III Limestone, MIS 7
Skeletal eolianites are superimposed on the partially calcitized old oolites and underlie the distinctive aragonitic oolite
of the last interglaciation. In most cases across the Bahamas,
only eolian facies are observed (e.g., Fig. 2, section 13), but in
New Providence Island, beach fenestrae indicate that sea level
was at or near the modern datum during a late middle Pleistocene interglacial highstand (Hearty and Kindler, 1997).
These deposits lie seaward of the higher, older oolitic ridge
associated with the 300,000-yr U-series age mentioned previously and underlie the last interglacial oolite of the Nassau
Ridge. This stratigraphic position is the basis for our correlation with MIS 7. Although some samples have low concentrations of amino acids, the well-preserved samples from several
Bahama Islands yielded A/I ratios averaging 0.55 0.03 (17)
(Table 1).
Couplet IV Limestone, MIS 5e
Exposed MIS 5e oolitic deposits are a key stratigraphic
marker because they are widespread and easily recognized
(Kindler and Hearty, 1996). They are limestones composed of
tangentialaragonitic ooids with thick corticies cemented by
low-Mg calcite equant spar. They comprise subtidal and beach
facies above present sea level, and their dunes form the highest
and most conspicuous ridges on most islands. Two distinct
phases of deposition are recognized in the last-interglacial

167

BAHAMAS STRATIGRAPHY AND SEA-LEVEL HISTORY

TABLE 1
Summary of Mean A/I Ratios (Aminozones) per Island and per Stratigraphic Unit from the Bahamas
(Listed by Decreasing Latitude) a
MIS
(Limestone
couplet)
Latest 1
(VI late)
Mid 1
(VI middle)
5a
(V)
5eII c
(IV late)
5eI c
(IV early)
7
(III)
9/11
(II)
13
(I)

Grand
Bahama

Abaco

Andros

New
Providence

Eleuthera

Cat

0.04
0.00(2)

0.36(1)
0.41
0.01(7)
0.47
0.02(3)

0.04(1)

0.12(1)

0.09(1)

0.28(1)

0.27
0.00(2)
0.36
0.02(5)

0.36
0.02(2)

0.37(1)

0.11
0.03(3)
0.29
0.03(5)
0.38
0.03(16)

0.12
0.01(2)
0.37
0.03(4)

0.47(1)
0.54
0.02(3)
0.64
0.08(2)

0.69(1)

Exuma

0.56
0.02(3)
0.69
0.03(4)

0.58
0.02(3)
0.66
0.05(16)
0.76
0.04(2)

0.57(1)
0.65(1)

0.10
0.02(4)
0.34
0.05(2)
0.40
0.03(25)
0.49
0.02(2)
0.59(1)
0.67
0.02(8)

San
Salvador b
0.09
0.02(6)
0.15
0.08(4)
0.33
0.04(7)
0.40
0.04(18)
0.50
0.06(18)

Long

0.10
0.01(3)
0.32
0.02(2)
0.43
0.02(5)
0.55(1)

Acklins Mayaguana

Caicos

Inagua

0.06
(1)

0.08
0.01(2)

0.05
0.01(2)

0.45
0.01(2)

0.42
0.02(4)

0.48
0.01(3)

0.10
0.01(2)
0.44
0.06(4)
0.57
0.01(2)

0.69
0.01(3)

0.77(1)
0.87(1)

0.93(1)

A/I values for the Inter-Laboratory Comparison Standards ILC-A, ILC-B, and ILC-C measured at the Utah Amino Acid Laboratory during the course of
analyses reported in this study (19931997) were 0.155 0.005, 0.520 0.023, and 1.095 0.045, respectively, and 0.148 0.004, 0.498 0.022, and
1.049 0.025 Northern Arizona University (1998 1999) using the same equipment. These values are well within the range measured for the same samples by
other laboratories (Wehmiller, 1984).
b
San Salvador aminozone averages are the result of combined values from Hearty and Kindler (1993a) and this study. Although different methods and
analyzers were used, aminostratigraphic results are comparable but yield greater variance. The error resulting from replicate analyses of the same geological unit
is reported in all other tables and figures in the format of the mean (X ), 1 standard deviation (1), and the number of samples analyzed (in parentheses) (e.g.,
0.67 0.02(5)). The absence of a reported error (except the Appendix) indicates that only one sample was analyzed.
c
See text for explanation of stratigraphic evidence for two distinct depositional phases of MIS 5e.

stratigraphy (Hearty and Kindler, 1993a, b). The early phase is


identified as 5eI and the latter as 5eII.
Ninety A/I ratios of late MIS 5eII oolite from 75 sample sites
on 13 island groups yielded a regional mean A/I ratio of 0.40
0.03 which enable a correlation of deposits of this significant
sedimentary event across a warming gradient of 700 km between Abaco (A/I 0.36; 23C MAT) and Great Inagua
islands (A/I 0.48; 25C MAT) (Fig. 3). The low variance
associated with the Bahamas-wide 0.40 ratio from 90 samples
supports our position that reworking of older grains was not a
significant factor. An older A/I cluster (5eI) is also detected
on several islands (Table 1). The range of mean ratios within
MIS 5e is likely a function of 10,000 15,000 yr of prolonged
warmth during the interglaciation. The ages of MIS 5e deposits
have been confirmed by numerous uranium ages (e.g., Chen et
al., 1991).
Couplet V Limestone, MIS 5e and 5a
Late Pleistocene skeletal eolianites of post-5e age directly
overlie oolites of MIS 5e at several localities in northern
Eleuthera (Fig. 4). A moderately well-developed paleosol and
karst surface separates the two units. Whether superimposed on
(Fig. 4) or laterally juxtaposed with 5e oolites, the two units
yield distinct mean A/I ratios (Table 1). The presence of MIS
5a eolianites on many islands indicates that sea level at that
time resided within 0 to 5 m of the present datum.
A relative-age difference between MIS 5e and 5a (expressed

by ) is represented by a consistent A/I of 0.09 0.03 (6


cases, 98 samples) (Table 1). Because the rate of epimerization
in skeletal and oolitic grains is equivalent (see below), the
offset cannot be attributed to differences in sample composition. In Bermuda, the A/I for 5e5a deposits is 0.06 0.02
(Hearty et al., 1992). The larger A/I in the warmer climate of
the Bahamas can be attributed to more rapid epimerization and
greater separation of the means. Cooler regional temperatures
for the duration of MIS 5, similarly recorded in marine isotope
records, would account for the relatively small amount of
epimerization between MIS 5e and 5a, compared to that within
MIS 5e (see above). Uranium-series ages on corals from MIS
5e and 5a deposits in Bermuda confirm a 40,000-yr difference
(125,000 85,000 yr) in age between the units (Ludwig et al.,
1996). We conclude that the A/I of 0.09, and the soil and
karst development from the Bahamian sites, represent 40,000
yr between MIS 5e and 5a. MIS 5c deposits have not been
recognized above sea level in the Bahamas.
Couplet VI Limestone, MIS 1
The middle Holocene is represented in the Bahamas by at
least two stratigraphically and petrographically distinct episodes of sedimentation occurring between 5000 and 3000 yr
ago (Carew and Mylroie, 1987). The oldest unit is characterized by thinly coated ooids deposited in eolianite units dipping
below 5 m on the windward bank margins. Whole-rock A/I
ratios from oolitic deposits average 0.11 0.01 (7) (MIS 1

168

HEARTY AND KAUFMAN

FIG. 3. Regional correlation of mean whole-rock A/I ( 1 s.d.) ratios (Table 1) with marine isotope stages and substages 1 through 9/11 plotted against
latitude. A/I ratios from the late last interglaciation (sensu stricto), MIS 5eII, forms the backbone of the aminostratigraphy. An early last interglacial aminozone
is represented by MIS 5eI. Mean A/I ratios generally increase with increasing abient temperature at lower latitude. The lack of trends in Holocene and modern
samples reflects their nearly unaltered state.

Oos, Appendix). An abrupt change in the depositional environment, perhaps corresponding with the growth of barrier
reefs up to the present sea level and diminished circulation in
the lagoon, initiated an environmental transition that favored
the formation of skeletal sediments about 3000 yr ago. A/I
ratios average 0.09 0.01 (6) from stratigraphically younger
skeletal Holocene units (MIS 1 Sks in Appendix) which
contain sea-level facies equivalent to the present datum.
Lightly cemented to unconsolidated subrecent beach and dune
samples (Mayaguana, Caicos, Inagua) yield mean ratios between 0.08 and 0.05, while active ooid shoals in Joulters Cays,
North Andros and Osprey Cay, Exuma Cays yield low A/I
ratios averaging 0.04 0.01 (3), indicative of active grain
formation over the past millennium.

SUMMARY AND DISCUSSION OF AAR RESULTS

Bahamas-Wide Concordance of Whole-Rock A/I Ratios


In a setting of considerable regional climatic variation (3C
difference in MAT and a precipitation gradient from 150 cm/yr
in the northern islands to 50 cm/yr in the southernmost islands)
and other local influences, the A/I data set nonetheless clearly
shows internal reproducibility, regional consistency, and maintenance of stratigraphic order in a majority of cases. A frequency histogram of A/I ratios (Fig. 5) summarizes the wholerock data from the Bahamas (Appendix) and identifies clusters
or aminozones named A, C, E, F, G/H, and I.
Each aminozone represents an interval of marine submergence and sediment production, while gaps between clusters

FIG. 4. Stratigraphic sections from Eleuthera showing skeletal eolianites from MIS 5a (couplet V) superimposed on oolitic deposits from MIS 5e (couplet
IV). The well-developed soil between couplets IV and V support the 30,000- to 50,000-yr hiatus indicated by the A/I ratios and radiometric dates. Data listed
in Table 1. Legend as in Fig. 2A. Site numbers in boxes appear in the Appendix.

BAHAMAS STRATIGRAPHY AND SEA-LEVEL HISTORY

FIG. 5. Frequency histogram of 227 whole-rock A/I ratios determined


during this study. Numbers above modes in distribution are correlated
oxygen-isotope stage. Capital letters are aminozones. Vertical arrows indicate the position of prominent red soils and stratigraphic and petrologic
features that separate aminozones (see text for discussion). Aminozones B
and D are reserved in the event that other interglacial deposits might be
discovered.

reflect intervals of emergence and nondeposition (i.e.,


limestonesoil couplets). Mean A/I values per lithostratigraphic unit per island group show that equivalent units on
different islands produce equivalent ratios (Table 1; Fig. 3).
Given the stratigraphic sequence, the aminozones reflect corresponding, distinct ages. Red soils (symbolized by arrows in
Fig. 5) identify approximate stratigraphic boundaries between
aminozones.
Because depositional units of equivalent age across the
region have similar compositions (Kindler and Hearty, 1996),
the potential variation in A/I ratios due to compositional differences within an interglacial unit is avoided or minimized.
Because most of the samples were extracted from aggradational coastlines (those with a positive sediment budget), mixing with older deposits is also expected to be minimal. However, if mixing of some grains from older interglacial deposits
did occur, the potential influence on the aggregate A/I ratio is
limited because the amino acid concentration decreases exponentially in older rocks (Hearty et al., 1992, and this study).
Thus, the high concentration of amino acids from the younger
grains overwhelms the potential influence of older reworked
grains. Our data in Table 1 are an empirical demonstration that
mixing of grains, should it occur between interglacial deposits,
is not sufficient to alter A/I ratios significantly.
Integrity of the AAR Whole-Rock Method
The whole-rock AAR application is a viable chronological
tool for regional correlation of limestone units. First, A/I ratios
correspond with the stratigraphic succession and vary in accordance with the defined lithostratigraphy. A/I ratios increase
with greater stratigraphic age, and deposits from different sites
that were previously correlated on the basis of independent
stratigraphic criteria, as discussed above, produce concordant

169

A/I ratios. A/I ratios in samples from coeval deposits across a


warming gradient increase with temperature (Fig. 3), as expected for this chemical reaction. The exception is Holocene
and modern samples that show only minor changes in amino
acid concentrations and A/I ratios compared with the initial
conditions.
Second, amino acid concentration systematically decreases
in samples of increasing stratigraphic age (Fig. 6). However,
because of decreasing concentration with age, older middle
Pleistocene and early Pleistocene samples are generally beyond
the effective range of the whole-rock AAR method and account
for the majority of the unexpected results. Thus, the practical
age limit of the whole-rock method in the Bahamas is approximately 450,000 500,000 yr (MIS 13).
Finally, we have established that the rate of isoleucine
epimerization in skeletal and oolitic limestones is indistinguishable. Heating experiments on Holocene samples demonstrate that skeletal and oolitic grains epimerize at nearly identical rates (Fig. 7A), while naturally epimerizing skeletal and
oolitic samples follow a parallel kinetic trend (Fig. 7B) over
the past half million years. In Bermuda and Hawaii, where all
deposits are skeletal, epimerization trends are parallel to those
from the alternating skeletal and oolitic deposits in the Bahamas (Hearty et al., 2000).
CALIBRATED AGES OF AMINOZONES

In previous paragraphs, we have demonstrated that wholerock A/I ratios are consistent with the relative-age sequence of
limestonesoil couplets. Independent radiometric ages generally support the whole-rock aminostratigraphy.
Because of their proximity and similar climate histories, A/I
ratios from New Providence and Eleuthera islands were combined to calculate a mean ratio of each aminozone. These
average values were then used to calculate calibrated ages
using the apparent parabolic kinetic (APK) model described by
Mitterer and Kriausakul (1989). The APK model is based on
the empirical observation that the rate of isoleucine epimerization in biocarbonates, held under isothermal conditions, decreases with increasing age as a parabolic function of time. For
this study, the APK rate constant is calibrated using the mean
A/I ratio (0.37) measured in MIS 5e samples and the median
accepted age of the interglaciation (125,000 yr) (Fig. 7B,
dashed line). Assuming an initial (t 0) A/I ratio of 0.01 to
account for the small amount of epimerization induced during
laboratory hydrolysis, as is typical for biocarbonates, the age
equation becomes
t 982.08 A/I 9.82 2 ,

(1)

where t time in years.


Based on the APK model, the mean calibrated age for each
aminozone falls within the range of odd-numbered MIS target
ages (Imbrie et al., 1984) for all but the MIS 7 unit, which

170

HEARTY AND KAUFMAN

duration and amplitude. The longest and highest highstands


would have the greatest impact on island growth. Several
depositional events stand out. An enormous volume of sedi-

FIG. 6. (A) Mean concentration of glutamic acid [Glu] in whole-rock


oolitic and skeletal samples analyzed in this study. Ages are inferred from
stratigraphic evidence and correlated with the global marine oxygen-isotope
chronology (isotope stage shown along side plot symbol). The decrease in
amino acid concentration with increasing age is fitted to a logarithmic curve.
Decreasing concentration rates with time are similar in both skeletal and oolitic
samples. (B) The exponential decrease in concentration is correlated with a
similar decrease in isoleucine epimerization rate, yielding an effective range of
the method of approximately 500,000 yr.

yields an older age (Table 2). We correlate aminozones A, C,


E, F/G, H, and I with MIS 1, 5a, 5e, 7/9, 11, and 13,
respectively. The APK ages of aminozones A, C, and E are
confirmed by stratigraphy and radiometric ages. The uncertainty in the ages of MIS 7 and 9 deposits might result from
changes in ambient temperature during the postdepositional
history of the samples. Alternatively, these deposits may have
originated during MIS 9, as suggested by the 300,000
22,000-yr calibrated age. If so, MIS 7 deposits may be scarce
or absent from the subaerial record of the Bahamas. MIS 11
(aminozone H) is associated with one of the longest and
warmest interglaciations of the past half million years, and its
APK age of 420,000 38,000 yr (Table 2) is tightly constrained by TIMS ages in Bermuda and Bahamas (Hearty et al.,
1999a).
IMPLICATIONS FOR ISLAND BUILDING
AND SEA-LEVEL HISTORY

The building of the Bahama Islands was modulated by


sea-level changes during interglacial highstands of variable

FIG. 7. Extent of isoleucine epimerization (A/I) measured in (A)


laboratory-heated, late Holocene whole-rock oolitic and skeletal samples
and (B) whole-rock samples from New Providence and Eleuthera Islands.
Data in both panels are plotted as the square root of time, as modeled by
apparent parabolic kinetics (APK; Mitterer and Kriausakul, 1989). (A)
Samples were moistened with purified water and heated in separate sealed
glass tubes at 140C. Numbers are heating times in days, error bars are 1
SD analytical errors of single heated sample, and lines are least-squares
linear regressions. The overlap in A/I measured in oolitic and skeletal
samples heated simultaneously indicates that these two biominerals epimerize at indistinguishable rates. The high r 2 values for the regressions (0.98)
show that the rate of epimerization conforms to a APK model over the
range of A/I values of interest. (B) The age of each unit is inferred from
stratigraphic evidence and correlated with the global marine oxygenisotope chronology (isotope stage shown next to plot symbol), error bars
are 1 SD of group mean A/I ratios based on multiple samples, summary
statistics and correlated ages are listed in Table 2, solid lines are leastsquares linear regressions (r 2 0.99), and dashed lines show the APK
model calibrated using MIS 5eI deposits alone (*). If the correlated ages are
correct, then the higher rate of epimerization shown by the regression
through all of the data, compared with the lower rate calculated using the
MIS 5e calibration, suggests that the middle Pleistocene samples experienced a somewhat higher average postdepositional temperature than the
samples deposited during late Pleistocene time.

BAHAMAS STRATIGRAPHY AND SEA-LEVEL HISTORY

TABLE 2
Estimates of Ages of Aminozones by Apparent Parabolic Kinetics (APK; Mitterer and Kriausakul, 1989) from Mean New Providence (NPI) and Eleuthera (ELU) A/I Ratios

Aminozone

Correlated
isotope
stage

A/I
NPI ELU
mean

A
C
E
F
G/H
I

1
5a
5e
7
9/11
13

0.11 0.01 (4)


0.29 0.02 (7)
0.37 0.02 (21)
0.57 0.02 (6)
0.67 0.03 (20)
0.76 0.04 (2)

APK
calibrated age
(10 3 yr)

MIS
age a
(10 3 yr)

9.6 1.0
76 10.8
125 b
300 21.6 c
420 38
542 58

010
7185
118128
186245
303423
478

Note. To estimate the uncertainty associated with the calibrated ages, we


apply the standard deviation measured for each aminozone (the betweensample variability) to the age equation derived from the APK model.
a
Ages of interglacial marine oxygen-isotopic MIS from Imbrie et al. (1984).
b
The age of 125,000 yr is used for calibration because it represents the
median age of the interglaciation, as does the A/I mean for 5e.
c
Because the APK model assumes a constant average temperature, an interval
of prolonged warmth up to the time of deposition of the calibration samples (i.e.,
MIS 5e) could have resulted in enhanced epimerization and an overestimate of the
ages of some of the older units. Thus, under natural temperature histories, epimerization may not strictly conform to a parabolic model.

171

ment was transferred from marginal shelves to the island core


primarily during MIS 11, and less so during MIS 9. These
events probably correspond with the Lower and Upper Town
Hill formations in Bermuda (Hearty et al., 1992, 1999a), and
the Kaena, Waialae(?), and Bellows Field events in Hawaii
(Stearns, 1978; Hearty et al., 1999b). Diminutive interglaciations during MIS 13(?) and 7 left little surficial record. Because
of voluminous sediment production, vast areas of the Bahama
Islands were buried in MIS 5e eolian and marine sediments.
During the present interglaciation, MIS 1, only minor island
growth has taken place.
Sea-level data from the Bahamas are summarized in Fig. 8
and Table 3. On tectonically stable carbonate platforms, accurate sea-level benchmarks (reflecting ocean volume) can be
established with fewer assumptions than on tectonic coastlines
or in deep-sea records.

CONCLUSIONS

(1) A succession of soillimestone couplets provides a


framework for testing the whole-rock method. We have determined that A/I ratios provide a reliable geochronological index
for Quaternary limestone units, and these units can be correlated across the Bahamas archipelago.

FIG. 8. Comparison of a normalized 18O curve (SPECMAP) for the past 500,000 yr (A) with a sea-level highstand reconstruction (B) based on geological
and geochronological studies in the Bahamas. Aminozones and limestonered soil couplets (IVI) are identified in the figure. Table 3 provides quantified
sea-level data.

172

HEARTY AND KAUFMAN

TABLE 3
Summary of Sea-Level Data from the Bahamas
over the Past Half Million Years
Marine
isotope
(sub)stage

Estimated sea level


(m present
datum)

Late 1
Mid 1
5a
Latest 5e

0 Datum
5
0 to 5
6 to 10

Early 5e

7
9
Late 11
Mid 11
Early 11
13

0 to 5
0 to 3
18 to 20
7.5
2
5 to 10

Limestone
couplet
VI
V
IV

III
II

Regional whole-rock
A/I ratio (Table 1
and [U-series])
0.09 0.01 (6)
0.11 0.01 (7)
0.31 0.02 (19)
0.40 0.03 (90)
[128-117 10 3 yr] a
0.48 0.04 (28)
[130-119 10 3 yr] b
0.55 0.03 (17)
[300 10 3 yr] a
0.68 0.03 (36)
[550 10 3 yr] c
0.76 0.04 (2)
[550 10 3 yr] c

Note. Representative stratigraphic sections, type localities, individual A/I


ratios, and references are provided.
a
Muhs et al., 1990.
b
Chen et al., 1991.
c
Hearty et al., 1999a.

(2) APK-calibrated age estimates show general correspondence with MIS 1, 5a, 5e, 9, 11, and 13. The age estimates
for MIS 7 are greater than indicated by the succession of
deposits, however, and appear to favor a correlation with MIS
9. MIS 7 deposits could be missing from the stratigraphic
record.
(3) An unambiguous correlation among 75 sites (90 ratios)
of late MIS 5e age on 13 islands forms the backbone of the
aminostratigraphy (Fig. 3). Extensive skeletal eolianites overlying and separated from MIS 5e deposits by a karst surface,
paleosol, and younger A/I ratios are correlated with MIS 5a,
about 80,000 yr ago.
(4) Sea level peaked at or well above present at least three
times during middle Pleistocene time since 450,000 yr ago. A
major drawdown of polar ice sheets occurred, probably during
MIS 11, when sea level approached 20 m above present. The
agreement of maximum highstand elevations during MIS 5e
and 11 support tectonic quiescence for Bermuda and the Bahamas.
(5) Identification of shoreline deposits representing at least
six interglaciations verifies that the depositional record of the
Bahamas is older and more complex than previously indicated
and that it has not been obscured by subsidence of the platform.
(6) We offer herein a first-order approximation of the ages
of several important interglacial highstands. Lacking evidence
of significant tectonic motion, the highstand deposits in the
Bahamas provide direct evidence for ice volume and global
climatic changes, as well as insights into the sedimentary

architecture and evolution of the Bahama Islands over the past


half million years.
ACKNOWLEDGMENTS
Research and logistical support in the Bahamas were provided by friends
and colleagues too numerous to mention. Reviewers comments were constructive and concise. J. Bright assisted in the Amino Acid Geochronology
Laboratory. This is a contribution to IGCP Project 437.

REFERENCES
Carew, J. L., and Mylroie, J. E. (1987). A refined geochronology for San
Salvador Island, Bahamas, In Proceedings of the 3rd Symposium on the
Geology of the Bahamas (H. A. Curran, Ed.), CCFL Bahamian Field
Station, pp. 35 44.
Carew, J. L., and Mylroie, J. E. (1995). Quaternary tectonic stability of the
Bahamian Archipelago: Evidence from fossil coral reefs and flank margin
caves. Quaternary Science Reviews 14, 145153.
Chen, J. H., Curran, H. A., White, B., and Wasserburg, G. J. (1991). Precise
chronology of the last interglacial period: 234 U/ 230 Th data from fossil
coral reefs in the Bahamas. Geological Society of America Bulletin 103,
8297.
Hearty, P. J. (1998). The geology of Eleuthera Island, Bahamas: A rosetta
stone of Quaternary stratigraphy and sea-level history. Quaternary Science
Reviews 17, 333355.
Hearty, P. J., and Kindler, P. (1993a). New perspectives on Bahamian geology:
San Salvador Island, Bahamas. Journal of Coastal Research 9, 577594.
Hearty, P. J., and Kindler, P. (1993b). An illustrated stratigraphy of the
Bahama Islands: In search of a common origin. Bahamas Journal of Science
1, 28 45.
Hearty, P. J., and Kindler, P. (1997). The stratigraphy and surficial geology of
New Providence and surrounding islands, Bahamas. Journal of Coastal
Research 13, 798 812.
Hearty, P. J., Vacher, H. L., and Mitterer, R. M. (1992). Aminostratigraphy and
ages of Pleistocene limestones of Bermuda. Geological Society of America
Bulletin 104, 471 480.
Hearty, P. J., Kindler, P., Cheng, H., and Edwards, R. L. (1999a). Evidence for
a 20 m middle Pleistocene sea-level highstand (Bermuda and Bahamas)
and partial collapse of Antarctic ice. Geology 27, 375378.
Hearty, P. J., Kindler, P., Cheng, H., and Edwards, R. L. (1999b). The Kaena
Highstand of Oahu Hawaii: Further Support for Partial Antarctic Ice
Collapse during Marine Isotope Stage 11, Fall Meeting 1999, American
Geophysical Union.
Hearty, P. J., Kaufman, D. S., Olson, S. L., and James, H. F. (2000). Stratigraphy and whole-rock amino acid geochronology of key Holocene and Last
Interglacial carbonate deposits in the Hawaiian Islands. Pacific Science 54,
4, in press.
Imbrie, J., et al. (9 authors) (1984). The orbital theory of Pleistocene climate:
Support from a revised chronology of the marine 18O record. In Milankovitch and Climate (A. L. Berger, et al., Eds.), Part I, pp. 269 305. Reidel,
Dordrecht.
Kindler, P., and Hearty, P. J. (1996). Carbonate petrology as an indicator of
climate and sea-level changes: New data from Bahamian Quaternary units.
Sedimentology 43, 381399.
Kindler, P., and Hearty, P. J. (1997). Geology of the Bahamas: Architecture of
Bahamian Islands. Developments in Sedimentology 54, 141160.
Ludwig, K. R., Muhs, D. R., Simmons, K. R., Halley, R. B., and Shinn, E. A.
(1996). Sea-level records at 80,000 from tectonically stable platforms:
Florida and Bermuda. Geology 24, 211214.

BAHAMAS STRATIGRAPHY AND SEA-LEVEL HISTORY


Miller, G. H., and Brigham-Grette, J. (1989). Amino acid geochronology: Resolution and precision in carbonate fossils. Quaternary International 1, 111128.
Mitterer, R. M. (1968). Amino-acid composition of organic matrix in calcareous oolites. Science 162, 1498 1499.
Mitterer, R. M., and Kriausakul, N. (1989). Calculation of amino acid racemization ages based on apparent parabolic kinetics. Quaternary Science
Reviews 8, 353357.
Muhs, D. H., Bush, C. A., Stewart, K. C., Rowland, T. R., and Crittenden,
R. C. (1990). Geochemical evidence of Saharan dust parent material for soils
developed on Quaternary limestones of Caribbean and western Atlantic
islands. Quaternary Research 33, 157177.

173

Muhs, D. R., Ludwig, K. R., Halley, R. B., and Shinn, E. A. (1994). Extended
Duration of Peak Last-Interglacial Sea-Level High-Stand from UraniumSeries Ages of Corals and Ooids from the Bahamas. AMQUA Program and
Abstracts of the 13th Biennial Meeting, 234.
Rutter, N. W., and Blackwell, B. (1995). Amino acid racemization dating.
Geological Association of Canada, Geotext 2, 125167.
Stearns, H. T. (1978). Quaternary shorelines in the Hawaiian Islands. Bernice
P. Bishop Museum Bulletin 237, 157.
Wehmiller, J. F. (1984). Interlaboratory comparison of amino acid enantiomeric ratios in fossil Pleistocene mollusks. Quaternary Research 22, 109
120.

Você também pode gostar