You are on page 1of 12

538 Chapter 1 5 1 Chronostratigraphy and Geologic llme

w E
· .
. : . .- . . ·. ·: : · ·. .- . . · . : : . . ·

1
. . · .. _ .: . · ·

Q)
E
i=

Figure 1 5.7
Transgressive-regressive cycle sedimentation and event correlation in the
Eocene of the Isle of Wight in southern England. [From Ager, D. V., 1 993, The
c::J CROSS - BEDDED SANDS

nature of the stratigraphical record, 3rd ed., Fig. 7.2, p. 1 00. Reprinted by § LAMINATED BEDS
permission of john Wiley & Sons Ltd. � GLAU CONITIC CLAYS

are related, resulting in a correlation in which glauconitic clays at the east end of the
succession are equated to laminated beds at the west end. Correlation is expressed,
as Ager (1993b, p. 101) puts it, in terms of degrees of "marineness." Correlation in
this manner can be considered to be a part of sequence stratigraphy (Chapter 13).

Correlation by Stable Isotope Events


Variations in the relative abundance of certain stable, nonradioactive isotopes in
marine sediments and fossils, referred to as stable isotope geochemistry (e.g., Val­
ley and Cole, 2001), can be used as a tool for chronostratigraphic correlation of
marine sediments. Geochemical evidence shows that the isotopic composition of
oxygen, carbon, sulfur, and strontium in the ocean has undergone large fluctua­
tions, or "excursions," in the geologic past-fluctuations that have been recorded
in marine sediments. Because the mixing time in the oceans is about 1000 years or
less, marine isotopic excursions are considered to be essentially isochronous
throughout the world. Variations in isotopic composition of sediments or fossils
allow geochemists to construct isotopic composition curves that can be used as
stratigraphic markers for correlation purposes. To be useful for correlation, fluctu­
ations in isotopic composition must be recognizable on a global scale and must be
of sufficiently short duration to show up as a shift on isotopic composition curves.
Also, stratigraphers must be able to fix the relative stratigraphic position of these
fluctuations in relation to biostratigraphic, paleomagnetic, or radiometric scales.
Of the various potentially useful isotopes, oxygen isotopes seem most nearly to
meet these requirements and have proven to be particularly useful for chronos­
tratigraphic correlation of Quaternary and late Tertiary sediments. Carbon, sulfur,
and strontium isotopes are also useful for correlating rocks of certain ages.

Oxygen Isotopes
The natural isotopes of oxygen are listed in Table 15.3. Most of the oxygen in the
oceans occurs as oxygen-16. Oxygen-18 is much rarer (about 0.2 percent of total
1 5.4 Chronocorrelation 539

Atomic Relative Isotopic


Element number Isotope abundances (%)
0 8 16 99.789

17 0.037

18 0.204

c 6 12 98.89

13 1.11

s 16 32 95.0
33 0.76

34 4.22

36 0.014

Sr 38 84 0.56

86 9.86

87 7.02

88 82.56

Source: CRC Handbook of Chemistry a n d Physics.

oxygen), but it is present in measurable amounts. The ratio of 180 / 160 in the
ocean at any given time in the past is built into contemporaneous marine carbon­
ate minerals and the calcium carbonate shells of marine organisms as a permanent
record of the isotopic composition of the ocean at those times. Fluctuations in oxy­
gen isotope ratios in the ocean with time thus show up in the geologic record as
fluctuations in the isotopic ratios of these marine carbonates and fossils. Classifi­
cation of deep-sea sediments on the basis of oxygen isotope ratios in the shells of
calcareous marine organisms, particularly foraminifers, has given rise to a new
stratigraphy for Quaternary sediments. This stratigraphic method is commonly
referred to as oxygen isotope stratigraphy. It was first used by Emiliani (1955),
who studied the isotopic composition of foraminifers in deep-sea cores and used
oxygen isotope ratios to subdivide the core sediments. Oxygen isotope stratigra­
phy has now developed into a major tool for correlating Quaternary and late Ter­
tiary marine successions, as explained below.
The 180/ 160 ratio in biogenic marine carbonates reflects both the tempera­
ture and the 180 / 160 ratio of the water in which these carbonates formed. The re­
lationships of ocean paleotemperature (T) to oxygen isotopic composition has
been shown by Shackleton (1967) to be

T(0C) 16.9 - 4.38( de dw) + 0.10(de dw )2 (15.1)

where de the equilibrium oxygen isotope composition of calcite and dw oxygen =

isotope composition of the water from which the calcite was precipitated. The de
and dw notations refer not to the actual oxygen isotopic abundances in calcite and
water but to the per mil (parts per thousand) deviation of the 180/ 160 ratio in cal­
cite and water from that of an arbitrary standard. A commonly used standard for
oxygen isotopes in the past was the University of Chicago PDB standard, where
PDB refers to a particular fossil belemnite from the Pee Dee Formation of South
Carolina. More commonly now, the isotope composition of ocean water (Standard
540 Chapter 1 5 I Chronostratigraphy and Geologic Time

Mean Ocean Water, or SMOW) is used as a standard (e.g., Coplen, Kendall, and
Hopple, 1983). The per mil deviation from the standard, referred to as a 1so, is ex­
pressed by the relationship

[ ( 1S0/ 160 ) sample - es0/ 160) standard]


a 1so x 1000 (15.2)
es0/ 160) standard
=

Oxygen isotope stratigraphy is based on the fact that a1so values in biogenic ma­
rine carbonates reflect both the temperature and the isotopic composition of the
water from which the calcite precipitates. These factors are both, in tum, a func­
tion of the climate. When water evaporates at the surface of the ocean, the lighter
160 isotopes are preferentially removed in the water vapor, leaving the heavier
ISo in the ocean. This isotopic fractionation process thus causes water vapor to be
depleted of ISO with respect to the seawater from which it evaporates. When
water vapor condenses to form rain or snow, the water containing heav� oxygen
will tend to precipitate first, leaving the remaining vapor depleted in so com­
pared to the initial vapor. Thus, the precipitation that falls near the coast and runs
back quickly to the ocean will contain heavier oxygen than that which falls in the
interior of continents or in polar regions, where it returns more slowly to the
ocean. The 1 so; 160 ratio of precipitates also correlates with the air temperature.
The colder the air, the lighter the rain or snow (Odin, Renard, and Grazzini, 1982).
For example, the overall average oxygen isotope composition of seawater is
--0.28 %o (per mil); however, the precipitation that falls in the crests of the Green­
land Ice Sheet is about -35%o and in relatively inaccessible parts of the Antarctic
Ice Sheet it is as negative as -58 %o .
The 1 s0-depleted moisture that falls in polar regions is locked up as ice on
land and is thus prevented from quickly returning to the ocean. Because of this re­
tention of light-oxlsgen water in the ice caps, the ocean becomes progressively en­
riched in lSo as sO-depleted ice caps build up during a glacial stage. Marine
carbonates that precipitate in the ocean during a glacial stage, particularly bio­
genic carbonates such as foraminifers, will be enriched in 1s0 relative to those that
precipitate during times when the climate is warmer and ice caps are absent, or
are much smaller, on land. Changes in the a 1so content of biogenic marine calcite
thus reflect changes in the volumes of ice on land and concomitant changes in sea
leveL That is, sea level drops as ice masses build up on land during glacial e�isodes
and rises when continental ice masses melt during interglacial stages. The a 80 val­
ues of seawater track these changes, becoming higher (more positive values) dur­
ing glacial stages when heavy oxygen is concentrated in the ocean and lower (more
negative) during interglacial stages as melting of polar ice caps returns light-oxy­
gen water to the oceans. These principles are illustrated in Figure 15.8.
Decrease in temperature of the seawater in which biogenic calcite precipi­
tates also causes an increase in the a1so values that are built into the calcite. Thus,
during glacial periods both decrease in temperature of ocean water and changes
in isotopic composition of ocean water owing to buildup of ice caps on the conti­
nents combine to increase the a1so content of biogenic calcites. Conversely, melt­
ing of polar ice caps, with consequent return of light-oxygen water to the oceans,
and increase in ocean temperature will be reflected in a decrease in 81so values in
marine biogenic carbonates.
Different kinds of marine organisms tend to incorporate somewhat different
ratios of oxygen isotopes into their shells (fractionate oxygen isotopes to different
degrees), as indicated in Figure 15.9. Therefore, to evaluate changes in oxygen iso­
topes in the ocean as a function of time requires that we analyze the same kind of
fossil organism in rocks of different ages. Planktonic foraminifers are the most
common fossil used in oxygen isotope studies of this kind.
1 5.4 Chronocorrelation 541

d1 8Q (%o}
0 � 0.5 - 1 .0 - 1 . 5

-oE- Full I nterglacial: sea level high stand

Termination of glaciation: rapid isotopic


..,.._
and sea level change

Full Glacial: ice volume maximum, sea level miminum,


..,.._ 18
0 enrichment in oceans

Period of slowly increasing ice volume and


sea level lowering
\
t Figure 1 5.8
Schematic illustration of the relationship between
Q) Last Interglacial: arrows denote
E
sea level changes continental glaciation and the o 1 80 content of
i=
ocean water during the last 1 50,000 years. Note
that o1 80 values become less negative (more
heavy oxygen) during stages of continental glacia­
Termination of glaciation: rapid isotopic and
_.__ tion and lowered sea level. [After Williams, D. F.,
sea level change
I . Lerche, and W. Fult 1 988, Isotope chronos­
Full Glacial: - 1 50,000 years before present tratigraphy-Theory and methods: Academic
Press, Fig. 30 p. 58. Reproduced by permission.]
,

Green algae Shallow-water


marine
limestone

\ Deep-sea
limestone
Hermatypic
corals
Shallow-water
molluscs and
foraminifers

Freshwater
limestone
Figure 1 5.9
Distribution of o 1 80 and o1 3 C values in various
types of marine carbonates. (After Milliman, J . D.,
1 9 74, Marine carbonates. Fig. 1 9, p. 33. Reprint­
ed by permission of Springer-Verlag.]
542 Chapter 15 I Chronostratigraphy and Geologic Time

Oxygen isotopes in sediment cores of late Tertiary to Quaternary age show


numerous 8180 maxima and minima. Figure 15.10 shows an example of oxygen
isotope values in three Pleistocene cores from widespread localities in the Pacific
and Atlantic oceans (Wei, 1993). The numbers on the isotope curves are oxygen
isotope stage numbers (e.g., Kennett, 1982; Ruddiman et al., 1989; Raymo et al.,
1989). Where available, the magnetostratigraphic chrons are shown also. Once the
isotope stages in a core have been identified and numbered, they can be correlated

Site 647 Site 593


1 31 80
8 80
5 4 3 2 0 4 3 2 0

10

30
20 10

0 35
0
'iii
<J) 30
flJ 15
:::
0
a>
.0 40
flJ
m
a; 40
:2
20
45

50
25
50

Figure 1 5.10 60
Oxygen-isotope stratigraphy 55 30
of Pleistocene cores from
Deep Sea Drilling Project
(DSDP) sites in the Pacific and
Atlantic oceans. Note the
n umbered isotope stages,
which can be correlated from 60
one core to another. M/B
refers to the Matuyama/Brun- 40
hes polarity chrons. [After 20
Wei, W., 1 99 3, calibration of 0
upper Pliocene-lower Pleis- 20
tocene nannofossil events
40
with oxygen isotope stratigra-
phy: Paleoceanography, v. 8., 60
Fig. 4, p. 9 1 , published by the
American Geophysical Union.] 1 20 1 60E 1 60W 1 20 80 40 0 40 80
15.4 Chronocorrelation 543

to the same isotope stages in other cores across the world ocean. These isotopic
events appear to be related to the Milankovich orbital-climate cycles discussed in
Chapter 12. Thus, they provide a record of fourth- and fifth-order cycles that are
presumably driven by changes in climate related to changes in Earth's orbital para­
meters. These orbital changes cause fluctuations in the intensity of solar radiation
reaching Earth at different latitudes, which, in tum, result in alternate accumula­
tion and melting of continental ice sheets, producing rise and fall of sea level.

Carbon Isotopes
Carbon-12 and carbon-13 are the nonradioactive isotopes of carbon. Carbon-12 is
much more abundant than carbon-13 and makes up most of the carbon in seawater
(Table 15.3). The isotopic 13C/1 2C ratio can be expressed in terms of per mil deviation
( 8 13C ) from the PDB standard, just as oxygen isotope ratios are expressed. The 8 13C
values in marine carbonates reflect the 13C/12 C ratio of C02 dissolved in deep ocean
water; this ratio, in tum, reflects the source of carbon in C02 • Carbon dioxide dis­
solves in the ocean by interchange with the atmosphere, and it is generated also by the
decay of organic matter that originates both in the ocean and on land. Organisms pref­
erentially incorporate light carbon ( 1 2C ); therefore, carbon dioxide derived from de­
caying organic matter is sharply depleted of 13C compared to that derived from the
atmosphere. Thus, water runoff from the continents (where soil organic matter is
abundant) brings organic-rich waters with low 13C/12 C ratios into the ocean, signifi­
cantly lowering the 813C content of surface ocean waters near the continents. (Note
from Fig. 15.9 the low 8 13C values in freshwater carbonates deposited in lakes.)
Another factor that influences the 813C content of ocean water, and thus the 8 13C
content in the shells of marine organisms that live in these waters, is the residence
time of deep-water masses in the ocean. Carbon-13 is depleted in deep-water masses
that have long residence times near the ocean bottom, owing to oxidation of low-8 13C
marine organic matter that sinks from the surface. Oxidation of this low-a13C organic
matter leads to production of low-a13C dissolved bicarbonate ( HC03 -). Respiration
by bottom-dwelling organisms also aeparently causes a decrease in a 13C of deep bot­
tom waters (Kennett, 1982) . If low-aL C bottom water is later circulated to the surface
in some manner, carbonate-secreting organisms will build this low-a 13C isotope ratio
into their shells.
The a Bc content of ocean water is also related to the primary productivity of
photosynthesizing marine organisms (e.g., diatoms). During times of high pro­
ductivity (large numbers of organisms present), the rate of removal of light carbon
( 12C ) in C02 by these organisms is high. Selective removal of the light carbon
causes surface ocean waters to be relatively enriched in Be, resulting in positive
values of a 13C. The light carbon is delivered to the seafloor when organisms die,
where it is reoxidized (Holser and Magaritz, 1992). During times of low primary
productivity, this trend is reversed and surface waters tend to have negative a 13C
values. It has been suggested, for example, that the dramatic decrease in aBc val­
ues across the Cretaceous-Tertiary boundary (see Fig. 15.11) is the result of marked
decrease in marine photosynthesizing organisms owing to a catastrophic mete­
orite impact (e.g., Hsu et al., 1982).
Because the a Bc in the calcareous shells of marine organisms is a function of
the a c content of the waters in which they live, changes in the oBc content of
B
fossil marine organisms indicate changes in ocean water masses. Abrupt decreas­
es in the a 13c in fossil marine calcareous organisms may reflect changes in prima­
ry marine productivity, as suggested, or changes in deep ocean paleocirculation
and upwelling patterns that caused low-aBc deep waters to spread upward and
outward into other parts of the ocean. Or such decreases may reflect changes in
544 Chapter 1 5 I Chronostratigraphy and Geologic Time

o 1 3C (%o PDB)
Site 528 Site 525A

+1 +2 +3 +1 +2 +3 +4
44

48

52

-;::
56
_g_
<ll
Ol
<(
60

64

68

0 +1 +2 +3 +2 +3 +4
Site 527 Site 529
o 1 3C (%o PDB)

Figure 1 5.1 1
Carbon isotope stratigraphy of bulk carbonate sediments from Deep Sea Drilling Project
(DSDP) Sites 525A, 5 2 7, 528, and 529 in the Atlantic Ocean about 800 km off the coast of
Africa. The sites are about 40 to 70 km apart. [After Shackleton, N. ] ., and M . A. Hall, 1 984,
Carbon isotope data from Leg 74 sediments, in Moore, T. C., Jr., and P. D. Rabinowitz
et al., Initial Reports DSDP 74, Washington, U.S. Gov. Printing Office., Fig. 1 , p. 61 4.]

surface circulation patterns that brought low-8 13C surface ocean waters from con­
tinental margins into deeper basins. Significant increases in the total biomass pro­
duced on the continents during any particular geologic time interval could cause
an increase in runoff of low-813C to the oceans, an increase that may be reflected
also as an episode of low-8 13C surface water. Increased rates of erosion of organic­
rich sediments, such as dark shales and limestones, on land could produce much
the same effect of increasing runoff of low-813C in organic matter to the ocean.
These abrupt changes in circulation patterns or organic carbon runoff from the
continents may have affected the area of a single ocean basin, such as the Pacific,
or in some cases the entire ocean. On the other hand, increased rates of sediment
burial in the ocean may have the opposite effect of removing fine organic matter
containing low-813C from interaction with seawater. This removal would have the
effect of increasing o13C in ocean water. Several increases in 8 13C that took place in
the middle Miocene between about 12-18 Ma may be associated with exceptionally
rapid burial of organic carbon and lowered levels of atmospheric C02 (e.g.,
Woodruff and Savin, 1991).
Inasmuch as changes in the 8 13C content of the ocean are reflected in the 813C
content of marine calcareous organisms, these isotopic events or "excursions" can
be used for correlation, regardless of their exact causes. For example, several
major changes in the Tertiary 8 13C record include (1) a pronounced negative 813C
event across the Cretaceous-Tertiary boundary, (2) a positive event from the early
1 5.4 Chronocorrelation 545

Paleocene into the late Paleocene, and (3) a major a13C decrease across the Pale­
ocene-Eocene boundary (Fig. 15.11). There is also a gradual but significant a 13c
deterioration from the mid-Miocene to the Pleistocene (e.g., Williams, Lerche, and
Full, 1988, p. 54). Many smaller-scale fluctuations in the Tertiary a 13C record are
also present. These carbon isotopic excursions are essentially synchronous events
that can be correlated over wide areas of the ocean in DSDP and ODP cores.
Major carbon isotope events are also present in the older sedimentary record.
For example, large changes in carbon isotope content of carbonates have been
measured at the Precambrian/Cambrian boundary and the Permian/Triassic
boundary (Magaritz, 1991).

Sulfur Isotop es
Sulfur has four stable isotopes (Table 15.3); sulfur-32 is the most abundant, fol­
lowed by sulfur-34. The 34S/32S ratio is used in most stratigraphic studies involv­
ing sulfur isotopes and is expressed in terms of a34S, which is per mil deviation of
the 34S/ 32S ratio relative to a meteorite standard, troilite (a FeS mineral) from the
Canyon Diablo meteorite. Figure 15.12 shows the a34S values in various materials
relative to the Canyon Diablo standard (CDT).
The major means of sulfur isotope fractionation in the oceans is bacterial re­
duction of sulfate ( S04 2- ) in seawater to sulfides ( H 2S, Hs- , HS0 4 - ) . Bacterial re­
duction of dissolved seawater sulfate at the sediment-seawater interface causes
isotopic fractionation of the sulfate, thus enriching the remaining seawater sulfate
in a34S by about + 20 %o and depletion in the reduced sulfide by about -9%o (e.g.,
Schopf, 1980; Canfield, 2001). Precipitation of evaporites from d issolved marine
sulfates introduces an additional fractionation ( � + 1 .65 %o), causing the a34S of
evaporites to be higher than that of dissolved marine sulfates. Other minor factors
that can influence the a34 S content of seawater include oxidation of bacterial
which produces sulfates depleted of 34S relative to original sulfates, and local em­
anations of sulfate or sulfide through volcanic activity.
Marine sulfates in the present ocean have a mean a34S of about +21 %a ,
however, the a34S of ancient marine evaporites ranges from about + 10 to +30
(Fig. 15.13). On the basis of sulfur isotope ratios in ancient evaporite deposits, it
appears that the sulfur isotope ratios in the surface waters of the world ocean have
undergone major changes, or excursions, at various times. These major excursions
are characterized by sharp rises in a 34S in the surface waters of the world ocean fol­
lowed by significant drops. The early Paleozoic exhibits high a34S levels, indicative

Canyon Diablo meteorite


I (standard)
Marine sulfates, recent

Marine sulfates, ancient

� Seawater sulfate
(2j Rainwater sulfate
Figure 1 5.1 2
- Sedimentary sulfides
a 34 S values of mari n e sulfates shown rela­
tive to that of the Canyon Diablo mete­
� 40 � 30 � 20 � 10 0 10 20 30 40
orite. Values for seawater sulfate, rainwater
su lfate, and sedimenta ry su lfides are shown
for comparison. [Data from Degens, 1 965.]
546 Chapter 15 I Chronostratigraphy and Geologic Time

0 ,--,-----.---,

1 00

200

Cil
6
(])
300

400

Figure 1 5.1 3
S ulfur isotope age curve for Phanerozoic marine evaporites. Dashed 500
portions of the curve indicate lack of data. The dashed vertical l ine
indicates the o 34 S composition of marine evaporites in the modern
ocean . [After Holser, W. T., M. Magaritz, and j. Wright, 1 986, Chem­ u

ical and isotopic variations in the world ocean during Phanerozoic 600 -'---"
a..
,_ ' +-r,.-.-..--,--.-,--,-.,-
- -r-'r-r-r,..,--.-,--,-,-rl
-
time, in Walliser, 0. (ed.), Global bio-events, Lecture Notes in Earth +10 + 20 +30
Sciences, v. 8, Fig. 4, p. 69, Springer-Verlag.] 8 348, 0/00 COT

of a high net flux of sulfide from ocean to sediments; 8345 drops to a minimum in
the Permian, and it returns to intermediate levels in the Mesozoic (Holser, Maga­
ritz, and Wright, 1986). Many of these sulfur isotope excursions appear to have af­
fected the ocean worldwide.
Chemical events characterized by sharply increased 8345 may be caused by
catastrophic mixing of deep 345-rich brines with surface waters. Brines generated
by evaporite deposition are stored in deep basins. Underneath the brines, bacteri­
al reduction of sulfates to form pyrite builds up a store of brine heavy in sul­
fate. Catastrophic mixing of these 345--rich brines with surface waters, owing to
destruction of the storage basin by tectonism, causes a sharp rise in the 8345 of sur­
face ocean waters and consequently in the evaporite deposits formed from these
surface waters (e.g., Holser, 1977; 1984). Gradual decrease in the 8345 of surface
ocean waters with time after a catastrophic event is attributed to on-land erosion
of predominantly sulfide materials into the ocean in an amount that exceeds evap­
orite deposition (e.g., Claypool et al., 1980).
In any case, these sulfur isotope excursions con.<>titute a sulfur isotope age
curve because each catastrophic chemical event occurred within a very short in­
terval of geologic time. Each major event thus represents a synchronous strati­
graphic marker that can be correlated in marine evaporite deposits from one area
to another. Some of the events can be correlated on a global basis. Thus, they pro­
vide an important method for international chronostratigraphic correlation of
evaporite deposits, which commonly cannot be correlated by other means because
they do not contain fossils or other datable materials. On the other hand, only a
few well-defined, correlatable points are present on the sulfur-isotope curve; thus,
sulfur isotopes are less useful overall for correlation purposes than are oxygen and
carbon isotopes.
1 5.4 Chronocorrelation 547

Strontium Isotopes
Strontium has four principal isotopes (Table 15.3), of which 87Sr and 86Sr are of
concern here. The relative abundances shown in Table 15.3 are somewhat variable
because 87Sr is the daughter product of the radiogenic isotope 87Rb (Table 15.3), as
previously discussed. The present-day quantity of 87Sr is thus a fraction of the ini­
tial amount of 87Sr in Earth plus the amount of radiogenic 87Sr generated from
decay of 87Rb through time (Veizer, 1989).
The amount of strontium in the ocean is derived by three mechanisms and
from three sources: ( 1) leaching of strontium from basaltic rocks owing to hy­
drothermal circulation at mid-ocean ridges (see discussion in Chapter 1), (2)
weathering of continental rocks and delivery to the ocean in river water, and (3)
diffusion of strontium from carbonate sediments caused by recrystallization dur­
ing burial diagenesis (McArthur, 1998). Because the 87Sr/86Sr ratio in each of these
sources is different, different ratios of 87Sr /86Sr are thus furnished to the ocean.
Strontium in river runoff has the highest 87Sr/86Sr ratio; strontium derived by hy­
drothermal circulation has the lowest. Weathering and river-runoff furnish the
greatest quantity of strontium; hydrothermal circulation at ridge crests is second
in importance.
The strontium isotope ratio of ocean water is constant throughout the ocean
at any given time but has varied through time owing to variations in strontium
contributed by these three processes. Variations in mid-ocean ridge volcanism and
the effects of changing world climates on weathering and river flow have exerted
the most important controls. Strontium is removed from the oceans by coprecipi­
tation with calcium in calcium carbonate minerals. Therefore, analysis of marine
carbonates of various ages allows the strontium isotope composition of the ocean
through time to be determined.
Figure 15.14 shows variations of the 87Sr /86Sr ratio of Phanerozoic-age ma­
rine carbonates, reflecting variations in these isotope ratios in ocean water since
Precambrian time owing to changing proportions of strontium contributed to the
ocean from different sources. Note a general decrease in 87Sr / 86Sr ratios of the
ocean from Precambrian time until the Jurassic, with several pronounced excur­
sions of lowered ratios (e.g., Ordovician, Devonian, Mississippian, Permian/Trias­
sic). A general increase in the ratio since Jurassic time is evident from Figure 15.14.
Strontium isotope data can be reported in terms of deviation from a stan­
dard, as in the case of oxygen, carbon, and sulfur isotopes (McArthur, 1998; Veiz­
er, 1989); however, many research results appear to be reported simply as 87Sr/86Sr
ratios. Isotope ratios can be used to establish correlation by findin� horizons with­
in two different stratigraphic sections that have identical 87Sr I 6Sr ratios (Fig.
15.15). Because more than one point on the strontium-isotope curve (Fig. 15.14)
can have the same 87Sr/86Sr value, independent evidence must be available (e.g.,
fossils, magnetostratigraphic data, regional geologic age relationships) to show
that the two horizons are approximately of equivalent age. Once correlation has
been established, it is possible to determine the numerical age (in millions of
years) of the correlation points from the isotopic age curve shown in Figure 15.14
(Faure, 1982; McArthur, 1994, 1998).
Correlation by strontium isotopes has been applied with particularly good
results to some Tertiary formations. For example, Depaolo and Finger (1988) use
strontium isotopes to correlate marine sediments of the Miocene Monterey For­
mation of California at resolutions comparable to those of biostratigraphic meth­
ods. By correlating the 87Sr/86Sr ratios in these sediments with the strontium
isotope vs. time curve for seawater derived from Deep Sea Drilling Program
(DSDP) coreholes in the southwestern and central Pacific, these authors deter­
mined ages with resolutions ranging from <0.1 to 2.5 Ma. See Bralower et al.
548 Chapter 1 5 / Chronostratigraphy and Geo lo gic Time

87Sr/86 r
s
.7060 .7070 .7080 .7090 . 71 00

200

"W
o:!
(].)
>-
0
tJl
c 300

I
(].)
Cl
<(

400

Figure 15.14
Variations of 87 Sr/86S r of seawater d u ring Phanero­
zoic time. [After Veizer, J., 1 989, Strontium isotopes
in seawater through time: Annual Review Earth and
Planetary Sciences, v. 1 7, Fig. 9., p. 1 5 7. Reproduced
by permission.]

(1997) for an application to Cretaceous deep-sea sediments recovered during


DSDP and Ocean Drilling Program (ODP) coring.

Problems with Isotopic Chronocorrelatlon


The field of stable isotope geochemistry is very complex; many questions remain
with regard to the validity of observed oxygen, carbon, sulfur, and strontium iso­
tope excursions in the geologic record and use of these excursions in chronocorre­
lation. A discussion of these problems and the details of correlation by stable
isotopes is outside the scope of this book. Additional information on this subject is
available in several of the publications listed at the end of this chapter.
Further Reading 549

SECTION 1 STRATIGRAPHIC SECTION 2 Figure 1 5.1 5


LEVEL
Schematic ill ustration of cor­
meters measured
87 86
=,...,==--- -----r 0 from cliff top relation by Sr/ Sr ratios
between two widely sepa­
meters
rated hypothetical strati­
0
graphic sections. If the
sections can be established
to have approximately
equivalent ages on the basis
of fossils or other age data,
levels within the stratigraph­
60 ic sections that have identi­
87 86
ca1 Sr/ Sr ratios formed

50
at the same time and there­
80
fore correlate. [After
McArthur, J. M., 1 998,
60 100
Strontium isotope stratigra­
phy, in Doyle, P., and M. R.
70 Bennett (eds.), Unlocking
0. 7076 0. 7078 0. 7076 0. 7078
the stratigraphical record:
"sr!''"sr Adva nces in modern stratig­
raphy: John Wiley & Sons
c:::J Carbonates BNi$,2 Clays !Wl;'!Jll Sandstone '17 0 .!. . Fossils
Ltd., Fig. 8 . 3, p. 228.]

FURTH ER READING

Berggren, W. A., D. V. Kent, M-P. Aubry, and J. Hardenbol (eds.), Odin, G. S. (ed.), 1982, Numerical dating in stratigraphy: John
1 995, Geochronology, time scales and global stratigraphic Wiley & Sons, ;\lew York, Parts I and II, 1040 p.
correlation: SEPM Spec. Pub. 54, Society for Sedimentary Ge­ Shiki, T., S. K. Chough, and G. Einsele, 1 996, Marine sedimentary
ology, Tulsa, Okla; 386 p. events and their records: Sedimentary Geology, v. 1 04, 255 p.
Clauer, N., and S. Chaudhuri (eds.), 1992, Isotopic signatures and [Special Issue]
sedimentary records: Lecture Notes in Earth Sciences 43, Snelling, N. J. (ed.), 1985, The chronology of the geological
Springer-Verlag, Berlin, 529 p. record: Geol. Soc. Mem. 10, Blackwell, Oxford, 343 p.
Clifton, H. E . (ed.), 1988, Sedimentological consequences of con­ Valley, J. W., and D. R. Cole (eds.), 2001, Stable isotope geochem­
vulsive geologic events: Geol. Soc. America Spec. Paper 229, istry: Reviews in mineralogy and geochemistry, v. 43, Miner­
157 p. alogical Society of America, 662 p .
Dickin, A. P., 1 995, Radiogenic isotope geology: Cambridge Uni­ Vertes, A . , S . Nagy, and K. Silvegh (eds.), 1998, Nuclear methods
versity Press, Cambridge, 452 p. in mineralogy and geology: Techniques and applications:
Faure, G., 1 986, Principles of isotope geology, 2nd ed.: John Wiley Plenum Press, New York, 555 p.
& Sons, New York, 589 p. Walliser, 0. H. (ed.), 1996, Global events and global event stratig­
Geyh, M. A., and H. Schleicher, 1990, Absolute age determina­ raphy in the Phanerozoic: Springer-Verlag, Berlin, 333 p.
tion: Springer-Verlag, Berlin, 503 p. Williams, D. F., I. Lerche, and W. E. Full, 1988, Isotope chronos­
Harland, W. B., R. L. Armstrong, A. V. Cox, L. E. Craig, A. G. tratigraphy: Theory and methods: Academic Press, San
Smith, and D. G. Smith, 1990, A geologic time scale 1989, 2nd Diego, 345 p.
ed.: Cambridge University Press, Cambridge, 263 p .
Hoefs, J . , 1997, Stable isotope geochemistry, 4th ed.: Springer-Ver­
lag, Berlin, 201 p.

You might also like