estimation of completeness of stratigraphical sections using

15
Journal of the Geological Society, London, Vol. 147, 1990, pp. 471-485, 14 figs, 1 table. Printed in Northern Ireland Estimation of completeness of stratigraphical sections using empirical data and theoretical models PETER M. SADLER' & DAVID J. STRAUSS2 'Department of Earth Science, University of California, Riverside, CA 92521, USA 'Department of Statistics, University of California, Riverside, C A 92521, USA Abstrad: The completeness of a stratigraphical section is the fraction of time intervals of some specified length (t) that have left a record. A record is left when some sediment is deposited during the interval and is not subsequently eroded. A complete section contains no hiatuses longer than t. The completenessof a section varies with t and its accumulation rate varies with the length of the time span over which it is measured. Plots of measured accumulation rate against time span are an empirical means of estimating completeness. Simple theoretical models help extrapolate meagre data and identify bias. Completeness at time scale t can depend upon three general properties of the accumulation history: the age of the section, the long term net accumulation rate (drift) and the unsteadiness of the sedimentation rate. The way unsteadiness is measured depends upon the kinds of fluctuations that can be recognized in the accumulation rate. To describe random fluctuations a standard deviation of rates is sufficient. One dimensional Brownian motion is a model of random fluctuations that explains many aspects of stratigraphicalcompleteness.Regularperiodicfluctuations in accumulation rate may be described in terms of wavelength and amplitude. Completeness is an increasing function of c, drift and, in the periodic case, the wavelength of fluctuations; it decreases with increasing standard deviation of accumulation rate andtheamplitude of periodicfluctuations.Thecompletenessand thickness of stratigraphical sections are weakly positively associated. A sound stratigraphical interpretation should not exceed the resolution of the stratigraphical section. However, it is not enough just to know the duration of the shortest event that can be resolved in the section. To interpret a succession of events, we need to know how completely the level of resolution is sustained throughout the section. In this paper, we examine the tenet that the completeness of a stratigraphical section can be estimated by plotting its accumulation rates against the time span for which they were determined. The sample of measured rates is typically small and biased; but we show how simple theoretical models of accumulation history help to complete the plot. Completeness and time scales Stratigraphical sections are archives of local geological history. The records are layers of sedimentary rock that are filed away in sequence, but numbered on a scale of rock thickness, rather than time. As a result of interludes of erosion or non-deposition, stratigraphical sections incorpor- ate hiatuses; in the terms of our metaphor, some records are removed, some are never filed and the archives become incomplete.Consequently,everysectionposes three basic time-stratigraphical questions. How much time do the individual rock layers record? How much time elapsed during the accumulation of the whole section? How much of the latter time interval was spent depositing the rock layers? This thirdquestion enquires about the possible complete- ness of the record of events preserved in the section. The answer has been termed the stratigraphical completeness of the section (Sadler 1981). Completeness is an important measure of the quality of a stratigraphical section as a source of geological data. It ought to be estimated before a section is claimed to be a 471 proving ground for any hypothesis about patterns of change. However, the completeness of a section doesnot havea general-purpose value. Consider the loss of a few daily records from historical archives; this does not diminish the value to the historian who seeks to reconstruct annual trends. Similarly, we evaluate the completeness of a stratigraphical section according to the precision required to test a hypothesis. Stratigraphical completeness must be defined and estimated with care. The next part of the paper shows that no value can be assigned to completeness unless we specify the time unit, or scale, of interest. This quantity, which we denote by t, plays a central role in the paper. The third part of the paper examines a method that estimates completeness from the fact that measured rates of accumulation also depend upon t. The method will prove to run foul of several sources of bias in empirical rate data and to require more dated horizons than most stratigraphical sections provide. For most stratigraphic questions, we would specify a value of t longer than the time scales at which anyone has observed sedimentary processes. Some modem sites of deposition have been monitored continually, but only for time spans that fall far short of the time it would take to accumulate a typical stratigraphicalsection.Consequently, when estimating completeness, we must be wary of the popular maxim that stratigraphical questions can be answered from experience in modern depositional environments. In the fourth part, we turn to simple theoretical models of sediment accumulation. They serve to quantify the controls of accumulation rate and completeness, identify the bias in empirical data and guide extrapolations from limited data.The models effectively bridge thegap between the time scales at which we observe sedimentological processes at Tulane University on September 27, 2012 http://jgs.lyellcollection.org/ Downloaded from

Upload: phamkiet

Post on 13-Feb-2017

215 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Estimation of completeness of stratigraphical sections using

Journal of the Geological Society, London, Vol. 147, 1990, pp. 471-485, 14 figs, 1 table. Printed in Northern Ireland

Estimation of completeness of stratigraphical sections using empirical data and theoretical models

PETER M . SADLER' & DAVID J . STRAUSS2 'Department of Earth Science, University of California, Riverside, CA 92521, USA

'Department of Statistics, University of California, Riverside, C A 92521, USA

Abstrad: The completeness of a stratigraphical section is the fraction of time intervals of some specified length ( t ) that have left a record. A record is left when some sediment is deposited during the interval and is not subsequently eroded. A complete section contains no hiatuses longer than t . The completeness of a section varies with t and its accumulation rate varies with the length of the time span over which it is measured. Plots of measured accumulation rate against time span are an empirical means of estimating completeness. Simple theoretical models help extrapolate meagre data and identify bias.

Completeness at time scale t can depend upon three general properties of the accumulation history: the age of the section, the long term net accumulation rate (drift) and the unsteadiness of the sedimentation rate. The way unsteadiness is measured depends upon the kinds of fluctuations that can be recognized in the accumulation rate. To describe random fluctuations a standard deviation of rates is sufficient. One dimensional Brownian motion is a model of random fluctuations that explains many aspects of stratigraphical completeness. Regular periodic fluctuations in accumulation rate may be described in terms of wavelength and amplitude. Completeness is an increasing function of c, drift and, in the periodic case, the wavelength of fluctuations; it decreases with increasing standard deviation of accumulation rate and the amplitude of periodic fluctuations. The completeness and thickness of stratigraphical sections are weakly positively associated.

A sound stratigraphical interpretation should not exceed the resolution of the stratigraphical section. However, it is not enough just to know the duration of the shortest event that can be resolved in the section. To interpret a succession of events, we need to know how completely the level of resolution is sustained throughout the section. In this paper, we examine the tenet that the completeness of a stratigraphical section can be estimated by plotting its accumulation rates against the time span for which they were determined. The sample of measured rates is typically small and biased; but we show how simple theoretical models of accumulation history help to complete the plot.

Completeness and time scales Stratigraphical sections are archives of local geological history. The records are layers of sedimentary rock that are filed away in sequence, but numbered on a scale of rock thickness, rather than time. As a result of interludes of erosion or non-deposition, stratigraphical sections incorpor- ate hiatuses; in the terms of our metaphor, some records are removed, some are never filed and the archives become incomplete. Consequently, every section poses three basic time-stratigraphical questions. How much time do the individual rock layers record? How much time elapsed during the accumulation of the whole section? How much of the latter time interval was spent depositing the rock layers? This third question enquires about the possible complete- ness of the record of events preserved in the section. The answer has been termed the stratigraphical completeness of the section (Sadler 1981).

Completeness is an important measure of the quality of a stratigraphical section as a source of geological data. It ought to be estimated before a section is claimed to be a

471

proving ground for any hypothesis about patterns of change. However, the completeness of a section does not have a general-purpose value. Consider the loss of a few daily records from historical archives; this does not diminish the value to the historian who seeks to reconstruct annual trends. Similarly, we evaluate the completeness of a stratigraphical section according to the precision required to test a hypothesis.

Stratigraphical completeness must be defined and estimated with care. The next part of the paper shows that no value can be assigned to completeness unless we specify the time unit, or scale, of interest. This quantity, which we denote by t, plays a central role in the paper. The third part of the paper examines a method that estimates completeness from the fact that measured rates of accumulation also depend upon t. The method will prove to run foul of several sources of bias in empirical rate data and to require more dated horizons than most stratigraphical sections provide.

For most stratigraphic questions, we would specify a value of t longer than the time scales at which anyone has observed sedimentary processes. Some modem sites of deposition have been monitored continually, but only for time spans that fall far short of the time it would take to accumulate a typical stratigraphical section. Consequently, when estimating completeness, we must be wary of the popular maxim that stratigraphical questions can be answered from experience in modern depositional environments.

In the fourth part, we turn to simple theoretical models of sediment accumulation. They serve to quantify the controls of accumulation rate and completeness, identify the bias in empirical data and guide extrapolations from limited data. The models effectively bridge the gap between the time scales at which we observe sedimentological processes

at Tulane University on September 27, 2012http://jgs.lyellcollection.org/Downloaded from

Page 2: Estimation of completeness of stratigraphical sections using

472 P . M . S A D L E R & - D . J . STRAUSS

and the time scales of the basic stratigraphical questions. Short term observations of modern sedimentary processes will suggest suitable parameters for the models, which can then be tested by the matching of simulated data from long term stratigraphical sections.

Stratigraphic completeness We define the completeness of a stratigraphical section at time scale t as the fraction of the time intervals of length t, that have left a record. For some purposes it is necessary to consider all intervals of length t within the time span T of the whole section; but it is often permissible and more convenient to consider only the consecutive time intervals. A record is left when some sediment is deposited during the interval and is not subsequently eroded. In a complete section all constituent time intervals of length t have left some sediment. A complete section is not without hiatuses, it simply contains no hiatuses longer than t .

Notice that completeness has the character of a probability. It must take a value between one and zero. It is a ratio of time intervals of length t. Those that have left a record are expressed as a fraction of all the consecutive intervals that might have left a record. Thus, expected completeness corresponds to the probability that some sediment will be deposited during a time interval of length t and survive subsequent erosion.

To see that completeness must depend upon time scale, consider the case where t equals T. All sections contain some sediment and must, therefore, be complete at this scale; regardless of the hiatuses, completeness equals one. As one attempts to resolve the section into increasingly finer time intervals, however, some will inevitably fall within hiatuses and completeness will decrease. Completeness assumes its lowest value when only one time interval has left a record; its value is then t / T . The lowest value for completeness depends upon t because the number of consecutive time intervals increases with the specified precision.

For a stratigraphical section that includes any erosional intervals, the complete history of accumulation (Fig. l a ) cannot be recovered. Even if all the sediment horizons were radiometrically dated, the most we could reproduce would be a ‘staircase’ plot (Fig. lb) in which the preserved increments of sediment appear as ‘risers’, separated by horizontal ‘treads’ that correspond to hiatuses. The history of deposition and subsequent erosion during the time now represented by hiatuses is almost always completely lost. A small component is occasionally reconstructible from trace fossils (Wetzel & Aigner 1986). Distinctive soils or surface encrustations may form during intervals of non-deposition, but those that survive erosion are imprecise guides to the time involved (Retallack 1984).

The reader might still imagine that t could be dropped from the definition of completeness if the accumulation were precisely monitored. Why is completeness not simply the fraction of time represented by the risers in the staircase plot? One reason is the particulate nature of sediment. Discrete sediment particles effectively arrive or depart instantaneously at the top of the section. The rate of pure deposition is infinite and the cumulative deposition time of all the sediment particles is zero. When examined in great detail, the staircase is in fact built of many small vertical risers, which cannot represent time. Instead, we divide the

1.

~ Magnetic polarlty reversal -

FIVE TIME SCALES lstlppled unltS leave a record)

Fig. 1. (a) History of the accumulation of a hypothetical stratigraphic section. Such a time series would be determined by continuously monitoring the level of sedimentation. (b) A stratigraphical ‘staircase’ plot for the same section. This curve includes only information recoverable from the stratigraphical section itself.

Large circles are schematic enlargements to indicate that the plots include unsteadiness at finer scales than can be drawn. The thickness of the section at the time of a polarity reversal (filled circle in (a)) is contrasted with the coordinates of the same event as preserved in the resulting stratigraphical section (open circles in (a) and (b)). Intervals characterized by erosion, or eroded sediment in (a), plot as horizontal steps in (b); they do not leave a record. As the time scale is more finely divided, a smaller proportion (stippled) of time units leave a record in the form of a preserved increment of sediment in (b). Notice that in this simple representation the fraction of time units that leaves a record may vary with the origin of the interval scales; completeness should consider all possible origins.

time scale into units of length t and determine how many of them include at least one positive step. Of course, some of the steps will be too small to resolve by any stratigraphical dating techniques. This is a significant bias in stratigraphical data, which we discuss later.

Figure l b copies typical stratigraphical practice and ignores all treads in the staircase plot less than some small value of I, thus allowing sloping risers that have finite time span. But natural hiatuses vary greatly in length and real sections may be incomplete at stratigraphically large values of t.

Factors that effect completeness Many important aspects of stratigraphical completeness can be appreciated directly from its definition. They will be fully

at Tulane University on September 27, 2012http://jgs.lyellcollection.org/Downloaded from

Page 3: Estimation of completeness of stratigraphical sections using

COMPLETENESS OF STRATIGRAPHIC SECTIONS 473

quantified during the later discussion of models, but it is useful first to determine properties that must be independent of the model. In this section, we seek a set of factors sufficient to determine completeness. In the process of defining completeness, we saw that it cannot be separated from the time scale t. This is the one factor that may be varied at the stratigrapher’s discretion. The others are all fixed during the accumulation history.

The primary controls of the accumulation history are a complex of interrelated geomorphological, ecological, climatic, tectonic and other aspects of the site and time of accumulation. It is hopeless to attempt an exhaustive list; instead we turn to the general properties of the resulting accumulation history and show that at least three are needed to determine completeness: (i) long term net accumulation rate; (ii) unsteadiness of sedimentation rate; (iii) age of the section.

The first of these, which we will refer to by its mathematical term drift, is evidently positively associated with the chance of surviving erosion and leaving a record. The second, which we shall simply call the unsteadiness, is negatively associated if drift is constant. Unsteadiness is the variability of accumulation rate as a function of time; it is a much more complex property than age or drift. The way it is measured depends upon the patterns of variability. If the variability is random, a standard deviation of rates is sufficient. Regular periodic fluctuations are usually de- scribed in terms of wavelength and amplitude.

Finally, the age difference between the top and bottom of the section is negatively associated with expected completeness. To see this, note that the chances of survival for different parts of a section will not be the same; they must be a decreasing function of distance below the top, i.e. age. Certainly, the probability of being eroded is greatest at the top of a section, as erosion to any depth removes the top of the section. However, the probability of being preserved at a given depth below the top is a different matter; preservation requires survival at all shallower depths during the burial process. The deeper and older parts of a section have, therefore, survived a greater chance of erosion.

Estimation of completeness from measured accumulation rates In this section we show that measured accumulation rates are dependent upon t and that the dependency reflects the presence of hiatuses. We will go on to discuss the problems which arise when this dependency is used to estimate completeness from measured accumulation rates.

Relation of accumulation rates and time span To appreciate the dependence of measured accumulation rates upon t , consider the time represented by a given thickness of sedimentary rock. It is well known that we cannot simply measure modem deposition rates to estimate how long it took to accumulate an ancient stratigraphical section (Tipper 1987). Imagine the selection of measure- ment sites in coastal regions or stony deserts; at some sites the surface of the sediment would be lowered by erosion and at others it would not change. Evidently, the accumulation rate, which we need in order to estimate time from thickness, is likely to be an average of both periods of aggradation and interludes of degradation or no change.

The rates of sedimentation measured from sites of modern deposition cannot include intervals of erosion or non- deposition that are longer than the period of measurement, which must be a small fraction of the accumulation time of most ancient sections. Consequently, the measured modem rates are systematically faster than longer term accumulation rates calculated between radiometric dates in analogous stratigraphical sections (Newell 1972; McKee et al. 1983). In fact, sediment accumulation rates generally decrease as the time span of measurement is lengthened (Reineck 1960; Schindel 1980; Sadler 1981).

If deposition is steady, accumulation rates are constant and thus independent of time and time span. If deposition is unsteady but continuous (i.e. variable but always greater than zero), the average accumulation rate remains independent of time span, but the variance of accumulation rates increases with decreasing t. Both cases produce complete sections at all scales by definition.

The average and the variance of accumulation rates, calculated from an incomplete (real) section, both tend to increase with decreasing time span. Incomplete sections arise when non-deposition or erosion interrupt the deposition process. If one could monitor the whole history of accumulation of the section (Fig. la), the average rate would still be independent of time span, because the net thickness is fixed (McShea & Raup 1986). Although the rates monitored at the site of deposition would include positive and negative values, only positive accumulation rates are recoverable from dated horizons in the resulting stratigraphical section (the staircase plot, Fig. lb). This omission of negative and zero rates causes the average measurable rate to increase with the variance. A broader frequency distribution of rates has both a higher maximum rate (included in the average) and a higher proportion of rates that are below zero (unpreserved and not included).

We have seen that interludes of erosion and non- deposition cause both accumulation rates and completeness to be decreasing functions of time span. Not surprisingly, the two are simply related. For a single stratigraphical section, the ratio of the overall accumulation rate to the average rate at time span t can be seen to give the completeness (Sadler 1981). The overall rate is total thickness divided by T. T is just Nt, where N is T l t or the number of non-overlapping intervals of length t. The short term average is the total thickness divided by nt, where n is the number of intervals of length t that are recorded by some sediment. So the ratio simplifies to nlN, i.e. to stratigraphical completeness.

Even for the most intensively dated stratigraphical section, there will not be enough dated horizons t units apart to determine the required short-term average rate exactly. One compromise is to calculate rates between any pairs of dated horizons that are available and plot all these rates against the time spans for which they are calculated. If a smooth curve can be fitted to these data, it offers a graphical means of estimating the average rate at time span t. Such plots of measured accumulation rates against time span are conveniently made logarithmic to display the full variability of natural rates and a geological range of time spans (Fig. 2). On the logarithmic plots the slope of the curve should be zero for time scales at which the section is complete. Zero slope simply indicates that the average accumulation rate is independent of time span. At time scales for which the section is incomplete, the average accumulation rate falls as

at Tulane University on September 27, 2012http://jgs.lyellcollection.org/Downloaded from

Page 4: Estimation of completeness of stratigraphical sections using

474 P . M . SADLER & D . J . STRAUSS

log. TIMESPAN

Fig. 2. A plot of accumulation rate against time span calculated for the hypothetical stratigraphical staircase Fig. lb. The mean (circle) and range (vertical bar) of accumulation rates are shown for discrete intervals of time span.

t increases and the regression has negative slope. Steeper negative gradients mean bigger discrepancies between short-term and long-term accumulation rate and thus decreasing completeness. The theoretical limit of the least complete sections is a gradient of -1. Since we have plotted

a.

0

0

h *. f . . -2 '4

-3

a fraction (thickness/time) against its denominator, lines with gradient -1 are contours of constant thickness. In other words, a slope of -1 arises if thickness is independent of time.

We have given constraints for the slope of these plots, but this is still far from a theory of their shape. We will turn to models for that theory, but first let us examine the form of plots of empirical data.

Estimates of completeness that use the data from a single section It is instructive to examine accumulation rates, as a function of t , for those rare stratigraphical sections that have abundant dated horizons. We shall find these empirical data subject to bias that frustrates attempts to estimate completeness for small values of t.

The best data for our present purpose are the published estimates of age for almost one hundred long cores recovered from calcareous oozes and chalks by the Deep Sea Drilling Project (DSDP). Figure 3 gives examples of plots of accumulation rate against time span for these cores. A little thought reveals other factors besides completeness that influence the slope of these plots: compaction, benthic mixing and a variety of measurement problems.

Obviously, the measured rates should be corrected for the effects of compaction. Closely spaced dates, from which short-term rates may be calculated, are most frequent in the youngest, least compacted, portions of a core. The longest-term rates are inevitably reduced by the inclusion of compacted chalks. This bias tends to exaggerate the negative slopes of plots such as Fig. 3, but can be removed using published measurements of density or porosity (Anders et al. 1987). In practice, we decompact the thickness to match physical properties after 1 m of burial.

b.

0 0

t

4 5 6 7 8 4 5 6 7 8 TIME SPAN ( log,, yr 1

Fig. 3. Examples of rate-time span plots for cores in calcareous oozes and chalks. Median rates calculated for constant time span are plotted as filled circles where based on actual compacted thicknesses, as open circles after decom- paction according to wet bulk density and as open squares after decompaction according to porosity. Crosses: median rates calculated for constant thickness, without decompaction. For clarity, only the compacted data are connected by linear interpolation. (a) DSDP 158; data include measurement bias at time spans less than 0.5 Ma. (b) DSDP 403; section includes a large hiatus, which produces a characteristic step in the trend when median rates are determined for con- stant time span and noise when they are measured for constant thickness.

at Tulane University on September 27, 2012http://jgs.lyellcollection.org/Downloaded from

Page 5: Estimation of completeness of stratigraphical sections using

COMPLETENESS OF STRATIGRAPHIC SECTIONS

This is below the layer where intense benthic mixing causes unpredictable porosities and tends to homogenize the distribution of radiogenic isotopes (Berger & Killingly 1982). Since the difference between neighbouring radio- metric ages is reduced in the benthic mixing layer, artificially high short-term rates are calculated. Of course, all of the core has passed through the mixing layer, so the effect resembles a moving average of sediment properties that is usually at least 1Ocm wide. On the empirical plots, data to the left of the 10 cm thickness contour (or any better, ad hoc estimate of the thickness of the mixing layer) are probably biased.

A third bias results from the difficulties of measuring small differences in age. To determine a short-term rate, it is necessary to find a small difference in age between two dated horizons. First of all, the most precise stratigraphical dating techniques are applicable only to relatively young rocks. So time span and age are not independent for these data and we have already demonstrated that age influences completeness. Furthermore, short-term rates may require more closely spaced samples than can be taken. A lower limit on the spacing of samples is a lower limit on the rates that can be measured at a given time span. Such a sampling limit obviously leads to some over-estimation of average rates at all time scales but, as time span decreases, the limiting thickness remains constant, the lower limit of measurable rates rises and the over-estimation becomes more extreme.

Anders et al. (1987) seized upon this sampling bias. In order to detect its presence, they suggest an examination of expected rates as a function of thickness. For their data compilations, it is not easy to specify what should be expected from this exercise. However, for single sections, it is elementary. If a section is subdivided into small time units, not all time units will have a sedimentary record. If a section is subdivided according to thickness, all subdivisions will include sediment; there will be no zero rates and no scale dependence of the average rate. In theory, the expected accumulation rate of a section is equal to the drift for all thicknesses and yet in practice the average accumulation rates of real sections, measured for constant thickness, decrease at small thicknesses. This is the opposite of the bias seen in rates measured for small timespans. The reason is still the difficulty of measuring small changes in age, As the section is divided into thinner and thinner intervals, it becomes more difficult to date them all. A measurably large difference is obtained between the ages of the top and bottom of the interval only if the accumulation rate is slow enough. In other words, average rate for a given thickness falls below the theoretical value (drift) when the sample of rates is inadequate; the point of departure from theory can be used as one indication of the scale at which an empirical data set becomes inadequate.

We should also mention an observational bias in published accumulation rates. Geologists measure what interests them. High accumulation rates are often a result of exciting events and they produce more impressive stratigraphical sections.

The biases we have identified are most extreme for short-term rate data. Unfortunately, empirical plots do not tell us how to extrapolate from the reliable long-term data into the short-term. Worse still, most stratigraphical sections are far too sparsely dated to establish any average rates reliably. One compromise is to combine data from many similar sections.

475

Estimates of completeness according to the environment of deposition If rates of accumulation that have been determined for the same environment of deposition, but at many different localities, are considered together a strong dependence upon t still emerges (Reineck 1960; Schindel 1980; Sadler 1981). Figures 4 and 5 show the results of such compilations for abyssal calcareous oozes and fluvial sediments respectively. Not surprisingly, such diagrams confirm that completeness may be expected to vary with the environment of deposition. In the plot of fluvial accumulation rates, the negative trend is much steeper than is found for abyssal oozes. So it is reasonable that we typically expect fluvial sections to be less complete at small time intervals. Notice that properties of the accumulation history reflect different controls in different environments. The long-term ac- cumulation rate, or drift, represents subsidence for fluvial deposition and the balance of average organic productivity and dissolution for abyssal oozes.

These compilations are subject to all the sources of bias mentioned above. A crucial question about the usefulness of a compilation of all rates from the same environment is this: does it describe a hypothetical average section? The answer is no. First of all, the compilations include the section-to-section, or lateral, variability (non-uniformity) of accumulation rates in addition to the purely temporal unsteadiness that determines completeness in a single section. The variance of rates for a given t is obviously inflated by the aggregation. A second difference to consider is that these data are derived from sections with a wide range of long-term net accumulation rates. Sections with different characteristics do not contribute evenly to all parts

- 0 L

\ X

E E

0, -2 m 0 d

Y

-4 a CK

-6

C

2 4 6 8 TIMESPAN (log,, y r )

Fig. 4. Scatter plot of about 22 OOO accumulation rates determined for calcareous oozes and chalks. Open circles: median rate as a function of time span. Open squares: median rate as a function of thickness. The data do not include compaction corrections. The trends of open circles and squares diverge at short time spans because of inadequate sampling; the divergence starts close to the thickness of benthic mixing layers.

at Tulane University on September 27, 2012http://jgs.lyellcollection.org/Downloaded from

Page 6: Estimation of completeness of stratigraphical sections using

476 P . M . S A D L E R & D. J . STRAUSS

\ 4 . - ' 2 ' d ' i ' i ' 6 l ' E ' -

T IME SPAN ( log,, yr l

Fig. 5. Scatter plot of 5600 accumulation rates determined for fluvial sediments in modem basins and ancient stratigraphical sections. Open circles: median rate as a function of time span. Closed circles: median rate as a function of thickness. The data do not include compaction corrections. The trends of open circles and squares diverge at short time spans because of the under- representation of low rates. The heavy dashes outline the field of data points from Fig. 4.

of the data base. Low ratios of drift to unsteadiness, for example, produce mostly stratigraphical sections of short duration and contribute little to the long-term rate data.

Single stratigraphical sections must be either active or relict. When accumulation rates are compiled for a whole environment of deposition, data from both types of section are apt to be mixed. Short-term rates have been determined for many environments by sampling near the surface of an

a.

- L Ot 0 1

b.

-1 1 -2 I_

active stratigraphical section. Here we may be dealing with the part of the staircase plot (Fig. lb) that is younger than the last erosional period and thus identical to the accumulation history. In some sedimentological studies, the active surface changes are monitored and it is then possible to record levels of the surface that are subsequently eroded. In this fashion, rates may be measured within and between segments of the accumulation history that are not routinely preserved in the staircase plot of an ancient stratigraphical section. Conversely, remember that many 'dates' are estimated in relict stratigraphical sections for levels of change in the preserved fauna, flora or magnetic polarity. These changes may have taken place during the time span of a hiatus. The dates and their heights in the preserved stratigraphical section then fall on a tread in the staircase plot and are not the coordinates of a condition that ever existed during the history of accumulation (they lie on the curve l b but not la).

Faced with a stratigraphical section for which there are few dated horizons, it is nevertheless tempting to substitute average rates determined from similar sections. In some cases, completeness has been estimated by the ratio of the locally measured drift to an average of short-term rates compiled for the same environment (Sadler & Dingus 1982; Gingerich 1982; Dingus & Sadler 1982). The procedure clearly breaks down when applied to sections that are not of average thickness for their duration (Fig. 6). Consider completeness at the time scale of the whole section. It must be 1.0. However, for thicker-than-average sections the values that result from this procedure are greater than one and obviously impossible. For thinner-than-average sec- tions, we would get values less than one; these are not beyond the range of values that completeness may assume, but they are equally impossible at the time scale of the whole section. The procedure in effect assumes that completeness is proportional to thickness (Fig. 6). We now examine that assumption.

Are thicker sections more complete? A section that is thicker-than-average might contain more than the average number of short-term increments and be

-I -+W-*+ + + + ++ I -tl+k+

5 6 7 8 4 5 6 7 8 TIME SPAN ( log,, y r 1

Fig. 6. Accumulation rates from (a) a relatively slowly accumulated (DSDP 536) and, (b) a relatively rapidly accumulated (DSDP 231) section of calcareous ooze and chalk compared with the median rates for such sedi- ments. Crosses: median accumulation rates. The slopes of tie lines from the long-term drift to the median rates indicate the error in estimating com- pleteness for these sections using median short-term rates and the ad hoc long term rate. Most tie lines misrepresent the trends of the single section data. Furthermore, the dashed tie lines would generate meaningless numbers for com- pleteness; negative slopes steeper than - 1 imply that a section preserves less than one time unit and positive slopes imply a completeness greater than one.

at Tulane University on September 27, 2012http://jgs.lyellcollection.org/Downloaded from

Page 7: Estimation of completeness of stratigraphical sections using

COMPLETENESS OF STRATIGRAPHIC SECTIONS 477

Table 1. Contingency table for 90 cores in calcareous ooze and chalk

Thinner Thicker than the median for than the median for this environment of this environment of deposition deposition

More complete than the median for l5 39* this environment of deposition Less complete than the median for 27* 9 this environment of deposition

* Fit the assumption that thicker sections are more complete.

more complete, but it might merely contain thicker-than- average increments. We have evaluated this question for 90 well-dated cores in calcareous oozes and chalks for which the relationship of rate to time span can be established. The thickness and completeness of these 90 sections were compared to the median values, at the same time scale, estimated from 243 cores in calcareous oozes and 122 sections in chalks. The thicknesses were compared at the time span T of each core. Relative completeness was estimated using the slope of the plot of median rate against t ; steeper negative slopes result from bigger discrepancies between long- and short-term rates and indicate less complete sections.

The values in Table 1 show a significant positive association between relative thickness and completeness. However, relative thickness is not a good quantitative indicator of completeness. Condensed (relatively thin, yet relatively complete) cores are seen to be a common exception; thick, yet relatively incomplete sections appear to be less common. Similar relationships were found for 15 sections in Cenozoic alluvial basins south of the Himalayas, but the association of completeness and thickness is weaker there.

Theoretical models of stratigraphical sections From an appropriate model of sediment accumulation, we can predict the dependence of accumulation rate upon t and avoid the biases that plague the empirical data. Complete- ness can also be quantified in terms of the four factors recognized above: time span, age, long-term net accumula- tion rate (drift) and the unsteadiness of the sedimentation process. They become parameters in the model. The first three are straightforward; unsteadiness needs more care. We first attribute unsteadiness in sedimentation rates entirely to random fluctuations. Subsequently, the influence of regular periodic fluctuations and intervals of non- deposition is modelled.

The main aim of this discussion is to give a simple account of a powerful Brownian motion model for the completeness of stratigraphic sections. The derivation of the model relationships has been described elsewhere in rigorous mathematical language (Strauss & Sadler 1989). We begin with a much more limited, but familiar, stochastic model: coin tossing. It is sufficient to derive several important lessons about random processes and stratigraphi-

cal sections. For all the properties that concern us here, the Brownian motion can be pictured as resulting from a long series of coin tosses, in a sense to be clarified later.

Random fluctuations modelled by coin tossing A limited stochastic model of the growth of a stratigraphical section may be set up as follows. Toss a fair coin at regular intervals; on heads let the section aggrade by one unit and on tails let one unit be eroded. One might expect that such a model would fail to replicate long hiatuses; this is not so. One might expect that half the runs of such a model, in which deposition and erosion are balanced, would fail to leave a section; we show that this guess is also wrong.

Feller’s (1968, ch. 3) analysis of coin tossing games shows that the thickness of sediment accumulated by this model will be ‘subject to chance fluctuations of a totally unexpected character’. Since the chances of erosion and deposition are equal, one might feel that the net thickness should frequently return to its initial value. Similarly, it might seem reasonable to expect the proportion of the accumulation history spent at a level of net erosion to be about one half. Feller (1968) shows that both guesses are far from the most likely outcomes. The effect of the random fluctuations is that the top of the section rarely returns to its initial level. In fact, regardless of the length of the accumulation history, the two extreme cases-the entire history either never attains, or never leaves, a state of net deposition-are most likely (Feller 1968, p. 78-80).

Large fluctuations in the history of net thickness arise routinely in such a random model. A fluctuation that takes longer to return to the initial thickness is likely to have been farther away from the initial value. We can model hiatuses at all time spans and the longer hiatuses are likely to involve deeper erosion. It is not necessary to build a complex periodic model to mimic the aspect of real stratigraphical sections.

Now it is correct to expect that half of a large number of coin-tossing simulations will end in a state of net erosion. However, net erosion implies only that the surface of the sediment is left below its initial elevation; it does not mean that there will be no sedimentary record. Most often the maximum downcutting will occur before the end of the accumulation history. Even a partial filling of the cut leaves a stratigraphical section, and the final step for half of all coin tossing runs must be a unit of deposition (heads!). Similarly, deeply eroded canyons usually have at least veneer of fluvial sediment on their floors.

Figure 7a shows how the proportion of runs that leave a stratigraphical section increases with the length of the accumulation history. The smaller numbers of tosses may not model anything stratigraphically useful, but the reader can easily verify the results for short runs by writing out all possible sequences of heads and tails. The limiting case is an infinite number of tosses, which is certain to leave a stratigraphical section; this is an important result for the Brownian motion model discussed below. The average accumulation rate of the simulated sections decreases systematically with time span (Fig. 7c & d). This average accumulation rate, being the proportion of tosses that leave a record, is also the completeness at the time interval of the coin tossing. Notice that completeness is correctly treated as a property of the stratigraphical product, not the sedimentological process; expected completeness is not

at Tulane University on September 27, 2012http://jgs.lyellcollection.org/Downloaded from

Page 8: Estimation of completeness of stratigraphical sections using

478 P . M . S A D L E R & D . J . STRAUSS

SECTION == 1 unit thlck

0

Length of occurnulatlon history lnurnber of tosses1 1 10 100 1000

t *

o m-] Fig. 7. Stratigraphical sections simulated by the coin-tossing model. The probability of generating a stratigraphical section (a) or a stratigraphical section containing only one increment (b) is a function of the number of tosses. The average or expected accumulation rate may be calculated for the total number of tosses (c) or the number of tosses between the first and last that are preserved in the section (d); curve d corresponds to radiometric data; bio- and magnetostratigraphical data would fall between curves c and d.

averaged across all possible runs, but only those that leave a stratigraphic section.

In Fig. 7 we distinguish rates whose time span is given by the total length of the accumulation history (Fig. 7c) from those where the time span must be determined by dating the first and last preserved increment of sediment within the section (Fig. 7d). The former include any hiatuses at the top and bottom of the section and are thus slower. They

a.

-4

b.

correspond to empirical rates calculated from a basement surface of known age up to the modern surface. We showed earlier that 'dated' changes in remanent magnetization or fauna may fall into the time span of a hiatus. So empirical rates based on the total thickness of sediment in a magneto- or biostratigraphical unit, of known time span, may similarly include parts of any hiatuses at the top and bottom of the record of the zone. The higher rates, based on intrinsic evidence of age, correspond to rates determined between radiometrically dated horizons. By distinguishing the two types of rate determination, we highlight a subtle bias in the empirical rate date. To the extent that the calculation of shorter term rates makes more use of the modern depositional surface, they will appear somewhat slower on average. All the sources of bias listed in the introductory section have the opposite effect.

Notice from Figs 7c and 7d, that the coin tossing model predicts that the gradients of the regression of average rate upon time span will have a slope close to -0.5. For long runs, the slope is exactly -0.5 (Strauss & Sadler 1989, Eqn. 2.6). We have seen that empirical plots can have much steeper gradients.

Random fluctuations modelled by Bernoulli trials Before progressing to the Brownian motion model, let us briefly put the coin-tossing model into its more general context. Coin tossing is an example of successive Bernoulli trials. The selections are independent and they have only two possible outcomes, for which the probabilities are constant. The coin-tossing model is a special case in which the probabilities of erosion and deposition are each one half. In the general case, these probabilities are unequal. After a large number of trials the expected accumulation rate tends not to zero, but to a value that depends on the difference between the probability of deposition and the probability of erosion.

Figure 8 shows the dependence of rate upon time span for sample runs in which deposition was more likely than erosion.

0

-2

-4

C.

0 2 4 6 0 2 4 6 0 2 4 6 log. TIMESPAN

Fig. 8. Accumulation rates from sample runs of Bernoulli trials. (a) 0.55/0.45 (probability of deposition/probability of erosion), unit thickness = 0.1, long-term net accumulation rate = 0.01; (a) 0.6/0.4, unit thickness = 0.1, long-term net accumulation rate = 0.02; (c) 0.51/0.49, unit thickness = 1.0, long-term net accumulation rate = 0.02.

at Tulane University on September 27, 2012http://jgs.lyellcollection.org/Downloaded from

Page 9: Estimation of completeness of stratigraphical sections using

COMPLETENESS OF STRATIGRAPHIC SECTIONS 479

Random fluctuations modelled by Brownian motion

For a model of sediment accumulation the fixed size of time and thickness increments in Bernoulli trials is a serious limitation. Tipper (1983) used a stochastic model that selects from sediment increments whose thicknesses have a continuous probability distribution. However, the fixed, integer, time units are a serious impediment when analyzing completeness. Firstly, the discrete time steps impose an arbitrary limit to precision. This is regrettable since we need to analyze completeness for a complete range of time scales. Secondly, the variable step sizes allow a sediment increment to be partially eroded; what remains then represents only a fraction of the basic, discrete, time unit and this complicates the analysis. Finally, the size of the time steps introduces a parameter with no practical geological counterpart; faced with a real stratigraphical section we have no information about the step size of the model. It is much better to make both the thickness and time scales continuous.

If the thickness increments are made sufficiently small and closely spaced, the accumulation history appears smooth, rather than stepped; a good analogue for the stratigrapher's view of sediment accumulation. Both the time scale and the thickness scale become continuous in the limiting case where time steps and thickness increments tend to zero. This limit, if appropriately taken (Strauss & Sadler 1989, Eqn. 3.2), gives a model in which the events are not as tangible as a falling coin, but like many mathematical limits, the model offers a powerful means of analysing the net effect of many events. The model is well known as one-dimensional Brownian motion. Some important fea- tures of the limit can be made nearly tangible by reference to the coin-tossing model. Imagine plotting an accumulation history modelled by an extremely long series of coin tosses and then viewing the result from a great distance. The discontinuous character of the plot becomes less evident as the series gets longer and the viewpoint more remote; at the limit, every discernible segment of the plot, however short, is made up of an infinite number of steps.

For the full mathematical analysis of stratigraphical completeness for a Brownian motion model the reader is referred to Strauss & Sadler (1989). The qualitative form of their results (summarized in Fig. 9) will be explained here by rather elementary dimensional analysis (Massey 1970, ch. 8; Yalin 1972, ch. 3). In short, we merely assume that the right equations must include all the relevant variables, must balance dimensionally and must match our earlier findings for the extreme, but simple cases.

We have seen that completeness is a dimensionless ratio that depends upon four variables, t, age, drift and unsteadiness. The only dimensions involved are height and time. So, from the basic theorem of dimensional analysis, it must be possible to write an equation in which completeness (already dimensionless) is a function of two (four variables minus two dimensions) other dimensionless products. They are chosen as follows.

One dimensional Brownian motion can be fully described by its drift and standard deviation. As controls of stratigraphical completeness, these parameters correspond to the net accumulation rate (subsidence, organic produc- tivity, etc.) and the unsteadiness of the sedimentation process, respectively. Just as it might be possible to estimate the balance between subsidence and sedimentation pro- cesses, without knowing the absolute rates of either, it is

convenient to measure the drift in units of the standard deviation. However, the two parameters have different units in this model. To build a dimensionless product, we multiply their ratio by the square root of t; the resulting dimensionless product is the dimensionless drift. For steady sedimentation the dimensionless drift is infinite.

Dimensionless drift does not include age, which must, therefore, be built into the second dimensionless product. What is meant by a 'long' burial history or an 'old' horizon varies with the stratigraphical context, as does the time scale at which we attempt to resolve completeness. So we can usefully generalize time by measuring it in units oft. In other words, age is divided by t and the last dimensionless product is built; it is the dimensionless age of a sediment increment or the dimensionless duration of a whole section (1 plus the dimensionless age of the bottom increment).

For a time interval to be 'preserved' in a stratigraphical

a.

nh 0 5 10 15

DIMENSIONLESS AGE

b. m v) W z W l- W J a E 0 U

W

U l-

W a X W

n

DIMENSIONLESS -DRIFT

Fig. 9. Completeness according to the Brownian motion model (after Strauss & Sadler 1989). (a) the probability of preservation of a time interval as a function of its age relative to the top of the section. To convert from dimensionless age units, multiply by the time span of the intervals considered. (b) expected completeness as a function of dimensionless drift for long sections.

at Tulane University on September 27, 2012http://jgs.lyellcollection.org/Downloaded from

Page 10: Estimation of completeness of stratigraphical sections using

480 P . M. SADLER & D. J . STRAUSS

section, some sediment must remain at the end of the time interval and it must escape erosion throughout the remaining accumulation history. For the limiting case of the Brownian motion model, all time intervals include an infinite number of steps; from the coin-tossing model (Fig. 7b), we know that this will guarantee that sediment remains at the end of the time interval. So it is necessary only to determine the probability that the thickness of sediment that remains exceeds the maximum net erosion during the rest of the run. Strauss & Sadler (1989) derive a complex equation for the probability of preservation that integrates across all thicknesses, from zero to infinity, that might remain at the close of a time interval. We now show how the graphical form of their result (Fig. 9a) can be anticipated without calculus.

Recall that the probability of preservation, averaged over all time intervals of length t, is the expected Completeness. The probability of preservation is a dimensionless ratio that is also a function of dimensionless drift and dimensionless age. We can easily graph the relationships between three dimensionless products by letting two form orthogonal axes and plotting a family of curves to show the effect of the third. An earlier section discussed how the probability of preservation varied with age (position in the section), so these two factors will be convenient axes. This leaves a family of curves to represent a series of values of dimensionless drift. Next, we determine the general shape of the family of curves by considering two extreme cases.

The probability of preservation is evidently one at the close of the interval in question. This is true regardless of dimensionless drift, so all curves converge at the point where dimensionless age is zero and probability is one. This also follows from the realization that expected completeness must become one at the time scale of the whole section. Now consider a long section. Near the bottom of the section, the chance of erosion is relatively insensitive to small differences in depth, SO the probability of preservation becomes independent of age. The family of curves must become parallel with the dimensionless age axis at large values of dimensionless age. This condition will be slower to develop where burial is slower, i.e. for lower values of dimensionless drift. That is enough information to sketch the form of Fig. 9a.

Figure 9a shows that, if a stratigraphical section has a sufficiently long dimensionless duration, the probability of preservation varies with depth only in a small region near the top. If we ignore that region, the probability is constant and dimensionless age is no longer needed. The expected completeness of a stratigraphical section becomes the same as this now constant probability of preservation and depends solely upon dimensionless drift. This is equivalent to setting dimensionless age to infinity and must be represented by a single curve on a plot of completeness against dimensionless drift (Fig. 9b). Again the general form of the graph is predictable. Expected completeness must tend to zero as dimensionless drift tends to zero, i.e. drift tends to zero, t tends to zero or the unsteadiness becomes large. Sections are, of course, complete if sedimentation is steady and dimensionless drift tends to infinity.

Figure 9b shows that expected completeness is an increasing function of drift and a decreasing function of unsteadiness. However, given that we have a section, the model shows that expected thickness is an increasing

W

k LL

n W l- V W a X W

m 0 -

1 I 1 c -1 0 1 2

log. TIMESPAN

Fig. 10. Median accumulation rates as a function of time span, according to the Brownian motion model (after Strauss & Sadler 1989). a, standard deviation is 1/10 of the drift; b, standard deviation is equal to the drift, c, standard deviation is 10 times larger than drift.

function of both drift and unsteadiness (Strauss & Sadler 1989). An increase in unsteadiness extends the range of outcomes, but we do not recognize the negative and zero thicknesses (no section). Thus, the model answers a question posed earlier: thickness and completeness will be strongly associated only when unsteadiness is nearly constant from section to section. A weak association will generally be found.

Figure 10 shows the expected short-term accumulation rates as predicted by the Brownian motion model. The figure has a limiting slope of -0.5 as t tends to 0. For large t , the expected rate is the drift. Figure 11 shows the short-term accumulation rates given by sample runs of the Brownian motion model. Evidently, a purely stochastic model can mimic empirical findings for deposits, such as abyssal oozes, where rate-time span plots do not show steep slopes except as an obvious result of benthic mixing. Figure 9b should satisfactorily give the expected completeness of a calcareous ooze and chalk section, provided that the standard deviation can be estimated. This simple stochastic model cannot account for the steeper empirical plots of average short-term rate such as found for the fluvial environment. The steeper plots are not necessarily due solely to biased data. A more complex random model, or periodic fluctuations, could account for steeper plots. For a modified random process, we suggest adding rapid and relatively large depositional steps at random to the Brownian motion. The result is that some runs tend to steeper slopes at small values of t. Non-random, astronomically controlled, periodic fluctuations are a popular theme in stratigraphy (e.g. Schwarzacher 1987; Weedon 1989). The next section shows how periodicity affects accumulation rates and can lead to steeper plots at stratigraphically significant values of t.

Periodic fluctuations modelled by sine functions For a simple model of periodic accumulation, we use the sum of a sine function and a fixed drift. Unlike coin tossing, this model is deterministic and the pattern of the accumulation history does not vary from run to run. All the

at Tulane University on September 27, 2012http://jgs.lyellcollection.org/Downloaded from

Page 11: Estimation of completeness of stratigraphical sections using

COMPLETENESS OF STRATIGRAPHIC SECTIONS 481

2

0

W l-

m Q -2

ci, 0 d

-4

a.

. . - X

- . . . - . -

A b. 2-

0-

. _ . -

-2- -

I 2-

0 -

-2 -

-4 -

log. TIMESPAN

Fig. 11. Accumulation rates from sample runs that were set up like the Brownian motion model but, of course, could not include an infinite number of points. The drift is the same for all three runs (0.01) but standard deviations differ (a: 10; b: 0.1; c: 0.01).

fluctuations are identical and the height of the sediment surface can be predicted exactly for any point on the time scale. To determine completeness from this model, we must replace the standard deviation of the stochastic model by the amplitude and period (wavelength) of the sine function. Length and time are still the only relevant dimensions, so completeness for the periodic model must be described by three other dimensionless products. Firstly, we build a dimensionless duration as before. Next, a dimensionless period is obtained by expressing the wave period as a multiple of t. Although the wave form of the accumulation history is exactly determined in this model, the end of the section may be a partial cycle. The effect of a partial cycle on completeness may be ignored, if the section is long relative to the wave period or relative to t. This consideration involves the first two dimensionless products.

The steepness of the wave form is important. It is given by the product of the amplitude and the frequency (reciprocal of period), which has the units of a rate. Divide this rate by the drift to obtain a dimensionless steepness, which is the final dimensionless product. When this model has zero drift, each fluctuation deposits sediment and then erodes exactly the same amount, leaving no section at the end of the cycle. When drift is positive, the depositional phase of each fluctuation accumulates more sediment than is eroded by the succeeding erosional phase. Every cycle leaves an increment of sediment, whose the thickness is given by the product of the drift and wavelength. Where the drift is not fast enough to counteract the maximum erosion rate achieved by the waveform (dimensionless steepness greater than 0.00278 for rescaled sine waves), each cycle also generates a hiatus. For positive drift, the fraction of the period occupied by hiatus is an increasing function of dimensionless steepness.

The preceding paragraph might leave the impression that the completeness of a long section is simply a function of dimensionless steepness. A little thought shows that the

dimensionless period is still critical. With positive drift, no time interval longer than the wave period can fail to be recorded by at least one sediment increment. All sections must be complete at values of t greater than the wave period. In fact, once t exceeds the length of one hiatus, the average accumulation rate becomes constant; it is the drift (Fig. 12). The simple periodic model has only one size of hiatus. As t becomes small relative to the wave period and the hiatus length, few time intervals that have left a record will include any hiatus. The average accumulation rate tends to a constant value that exceeds the drift (Fig. 12).

Sections thicker than one sediment increment are incomplete at time scales shorter than one hiatus. So values of t have an intermediate range, which expands with dimensionless steepness, where accumulation rates fall systematically with increasing t. In this range, the maximum thickness is one sediment increment (drift X wavelength), but the time spans include hiatuses. It follows that, on a plot of accumulation rate against time span (e.g. Fig. 12), the maximum rates in this range lie along a line of slope -1. This line is the thickness contour for one sediment increment. Whether or not any measured rates fall in this range depends upon the dating technique. The trend emerges when a section is dated using the levels of magneto- or biostratigraphical events; this is because, as we have seen, these ages may fall in hiatuses. Radiometric dates of the sediment increments generate no rates for this range of time spans (Fig. 12a).

As a model of episodic sedimentation, the single sine wave is unrealistic. It does not capture the particulate character of sediment; nor is it an appropriate model for completeness in sections thinner than one increment. Nevertheless, by considering a single sine wave, it is easy to appreciate the effect of adding some periodic fluctuations to a stochastic model. It is clear that adding a periodic fluctuation to the drift in the stochastic model will enable it to mimic the steeper rate-time span plots that typify the

at Tulane University on September 27, 2012http://jgs.lyellcollection.org/Downloaded from

Page 12: Estimation of completeness of stratigraphical sections using

482 P. M . S A D L E R & D. J . S T R A U S S

c

- L

W l-

c Q -4

-6

a.

- .

0

-2

-4

- E

C .

0 2 2 0 2 4! 6 0 2 4 4 6 log. TIMESPAN

F&. 12. Accumulation rates produced by the sum of rescaled sine waves and a fixed, positive drift (0.001). (a) amplitude 10 OOO, wavelength (arrow) 36 OOO, dimensionless steepness 278; rates calculated as if based upon radiometrically datable horizons only. (b) parameters as in (a), but rates calculated as if including biostratigraphically dated horizons; such dates may lie in hiatuses and fill the gaps in plot (a). (c) amplitude 1, wavelength 3600, dimensionless steepness 0.278.

empirical data from depositional environments at or above wave base. Figure 13 illustrates sample runs of a model in which the drift is not fixed, but follows a sine function. The effects of such a combination of the random and periodic fluctuations would have to be described in terms of four dimensionless products, the three for the periodic model and the dimensionless drift.

Models with non-depositional interludes

The last element of unsteadiness we consider is non- deposition. The simplest model adds a proportion of non-depositional steps at random to a stochastic process. The proportion of non-depositional steps appears to be an additional dimensionless factor, but it can be handled simply

2

0 W l- < U

ci, -2 0 d

1 .

C.

0 2 L 6 0 2 4 6 0 2 4 6 log. TIMESPAN

Fig. W. Accumulation rates from sample runs of Brownian motion models in which the drift varies as a rescaled sine wave. (a) drift 0.1, standard deviation 10, amplitude 10, wavelength 360, dimensionless steepness 0.278. (b) parameters as in (a), except standard deviation is ten times smaller. (c) drift 10, standard deviation 100, amplitude 100, wavelength 3600, dimensionless steepness 0.00278. Breaks in the clouds of points arise because rates are calculated as if based upon radiometric dates only.

at Tulane University on September 27, 2012http://jgs.lyellcollection.org/Downloaded from

Page 13: Estimation of completeness of stratigraphical sections using

COMPLETENESS OF STRATIGRAPHIC SECTIONS 483

0

U t-

U Q -2

m d 0

-4 -4

b. A 0 -

-2 -

-4 -

0 2 4 6 0 2 4 6 0 2 4 6 log. TIMESPAN

Fig. 14. Effect of non-depositional interludes. (a) Bernoulli trials with probabilities 0.26 for deposition, 0.24 for erosion and 0.5 for non-deposition; unit thickness = 1.0; long term net accumulation rate = 0.02; compare with Figs 8b and 8c: Brownian motion with drift 0.01. (b) standard deviation 0.1 and 90% non-deposition. (c) standard deviation 0.01 and 65% non-deposition.

in our models by modification of the drift value. In sample runs (Fig. 14a) where the other steps are determined by Bernoulli trials and the relative frequency of erosion and deposition are unchanged, non-deposition simply reduces the difference between the probability of deposition and erosion. The result is a proportionally lower net accumulation rate. For Brownian motion models (Fig. 14b), the random addition of non-deposition alters both the drift and standard deviation and results in a net reduction in dimensionless drift. Allowing non-depositional interludes simply changes the values of parameters in the stochastic models. It does not otherwise change the behaviour of completeness.

Models with memory Probabilisitc models in which the steps are completely independent have been considered. However, this omits the important Markov models in which the outcome at each step depends partly on the outcome of one or more preceding steps (Schwarzacher 1975). Such processes may be considered to have a memory of a certain length; what they do next depends upon what they have done recently or what position they have reached. Such models are relevant, especially to sections built up of several lithologies and they have been used to analyse cyclothems. We have begun to pick apart fluvial sections to separate the properties of channel and overbank deposition. To simulate these sections properly, the model needs to remember whether the last step was in a channel or overbank setting. Brownian motion still describes the limit of coin-tossing models with memory added, but the dimensionless drift value is again affected. Models can be written with memory effects that dampen the fluctuations relative to simple Brownian motion.

Estimates of completeness that use empirical data and theoretical models If we wish to estimate the completeness of a single stratigraphical section, two different strategies may be

followed that combine empirical data and the results of the modelling exercise. The most appropriate parameters to describe the sedimentation history (from comparable modem environments where necessary) are chosen and the resulting model is relied upon to predict completeness. The other uses model predictions of the dependence of average accumulation rate upon time span to fit the most reasonable curve to limited empirical rate data.

Faced with a real stratigraphical section, the drift and age are usually easy to estimate. The time span t is, of course, selected solely at the stratigrapher's discretion. Unsteadiness parameters are more difficult. For periodic fluctuations the wavelengths may be taken from tidal or Milankovitch cycles (Weedon 1989), or chosen to match sea-level curves. The amplitude, however, is measured in sediment thickness and it must be estimated from sediment cyclothems. Note that we need the amplitude of the depositional part of the cycle before erosion. In a rigidly periodic model, all cyclothems but the youngest are potentially truncated. It follows that we should estimate cyclothem amplitude from near the top of an actively accumulating section. Alternatively, we might use the thickness of the most complete ancient cyclothem. For random fluctuations in sedimentation rate, we need to estimate the standard deviation. The best estimates can be made where a modern site of accumulation has been monitored continually. Regrettably, such data are rarely published.

Inman & Rusnak (1956) documented accumulation histories for shelf sand near La Jolla, Southern California, with about 40 measurements taken at irregular intervals over 2.5 years. This is a depositional setting where non-random fluctuations are likely. Our estimations of standard deviation for different portions of the history show immediately that the process is not Brownian. Neither is it simply periodic. The process includes a strong damping factor (memory) that severely limits its ability to fluctuate far from the drift (which is zero); such a sedimentation pattern is unlikely to accumulate a significant stratigraphical section.

at Tulane University on September 27, 2012http://jgs.lyellcollection.org/Downloaded from

Page 14: Estimation of completeness of stratigraphical sections using

484 P . M. S A D L E R & D . J . STRAUSS

Notice that by estimating unsteadiness from the accumulation history, we can test the appropriateness of different models. Strauss & Sadler (1989) describe several tests for the Brownian motion model. It is, of course, simpler to model the distribution of hiatuses (staircase plot) directly, rather than via the accumulation history (e.g. Plotnick 1986). We contend that the usefulness of such models will be relatively limited, since the parameters may not correspond to measurable sedimentological or strati- graphical properties.

The second strategy to estimate the completeness of a single section follows four basic steps: (i) make a plot of accumulation rate against time span using only estimates of age from the section in question, however sparse they may be; (ii) indicate areas of the plot that are obviously biased by benthic mixing, a fixed sample spacing, compaction, etc.; (iii) choose the most appropriate class of model according to the expected importance of periodic (e.g. Schwarzacher 1987) and random fluctuations; (iv) fit a curve to the data that passes through the measured drift (longest term accumulation rate), comes close to other reliable data points and has a slope that resembles the prediction of the general model.

Unless we believe that the accumulation process had some regular periodicity or was subject to sudden fluxes of sediment, the modelling exercise tells us to suspect bias where the short-term portion of the plots has a negative gradient of -0.5 or steeper. Of course, plots for single sections are individual sample runs and may depart from the expectation of a random process. For some sections, we may accept that real periodic fluctuations account for the steep parts of the empirical plot. The steep parts become more pronounced as dimensionless steepness increases; we can estimate amplitude, wavelength and drift accordingly. Clearly, the estimation of completeness for a single section involves subjective, stratigraphical expertise. Careful con- sideration of the models and the definition of completeness has identified flaws in our intuitive expertise.

Whether we work from sedimentation parameters or empirical accumulation rates, the estimation of stratigraphi- cal completeness is typically frustrated by lack of data. However, the modelling exercise shows that absolute values are not always needed. The results were described in terms of dimensionless products. Similarly, estimates of the ratios of controlling factors are sufficient to compute complete- ness. At the simplest level, the models isolate the influence of each factor upon completeness. If it is known how any of these factors vary between the available sections, the models can help choose the best section for a particular study.

Summary An estimate of the completeness of a stratigraphical section is one aspect of its resolution. It is good to know that a section can resolve an event of duration t. However, geologic history is a succession of events; we should go on to ask what fraction of the time intervals of length t have left a record in the section. A record is left when some sediment is deposited during the interval and is not subsequently eroded. The fraction in the question is stratigraphical completeness and it is meaningless without f. A complete section is not one without hiatuses, it simply contains no hiatuses longer than t.

When the accumulation rates of a stratigraphical section

are determined for intervals that include hiatuses, the average rates become a decreasing function of the time span of the interval. This provides an empirical basis for estimating completeness. The relationship of average accumulation rate to time span is different when rates are determined using bio- or magnetostratigraphical time scales, rather than local radiometric ages. The empirical data are also typically biased and insufficient, especially for small values of t. It is practically impossible to determine the completeness of a single section without an infusion of subjective expertise, but the influence of random fluctua- tions in sedimentation can be counter-intuitive. Because the time scales of stratigraphical questions rarely match the time scales of our experience with modern sediment accumula- tion, a simple uniformitarian approach can be misleading. Random and periodic models of sedimentation can improve our judgement and bridge the gap in time scales. The models provide a theoretical basis for extrapolating from limited empirical rate data and for recognizing biased data.

Completeness has the character of probability; it must take a value between one and zero and it is the fraction of successes (leaving a record) in all possible outcomes. Expected completeness corresponds to the probability that a time interval of length t will leave a record, averaged over all such intervals in a section. This formulation is the basis for stochastic models of completeness.

The primary controls of sediment accumulation history are too complex to list. However, completeness at time scale t can be modelled from three general properties of the resulting accumulation history: the age of the section, its long term net accumulation rate and the unsteadiness of sedimentation rate. The first two are usually simple to estimate. The measures of unsteadiness must be suited to the kinds of fluctuations that can be recognized. For random fluctuations, a standard deviation of rates is sufficient and one dimensional Brownian motion is a model that produces answers to many basic questions about completeness. Periodic fluctuations are described in terms of wavelength and amplitude. Completeness is an increasing function of time scale, drift and the wave length of fluctuations; it decreases with increasing standard deviation of accumula- tion rate and the amplitude of periodic fluctuations. The completeness and thickness of stratigraphical sections are weakly positively associated, but the association is strongest when unsteadiness varies little from section to section.

This paper was significantly improved by suggestions from R. Howarth, G . Weedon and L. Dingus, who read an earlier draft. The compilation of data illustrated in this paper was supported by grant EAR-8305914, from the National Science Foundation. The modelling is a contribution to the ‘Consortium to dissect the Precambrian-Cambrian transition’ (NSF grant EAR-8721192).

References ANDERS, M. H., KRUEGER, S. W. & SADLER, P. M. 1987. A new look at

sedimentation rates and the completeness of the stratigraphic record. Journal of Geology, 95, 1-14.

BERGER, W. H. & KILLINGLY, J . S. 1982. Box cores from the equatorial

45, 93-125. Pacific: C-l4 sedimentation rates and benthic mixing. Marine Geology,

DINGUS, L. W. & SADLER, P. M. 1982. The effects of stratigraphic completeness on estimates of evolutionary rates. Systemtic Zoology, 31, 400-412.

FELLER, W. 1968. An Introduction to Probability Theory and Its Applications, Volume 1. Wiley.

at Tulane University on September 27, 2012http://jgs.lyellcollection.org/Downloaded from

Page 15: Estimation of completeness of stratigraphical sections using

COMPLETENESS OF STRATIGRAPHIC SECTIONS 485

GINGERICH, P. D. 1982. Time resolution in mammalian evolution: sampling,

Paleontolog@al Convention, 1, 205-210. lineages, and fauna1 turnover. Proceedings of the 3rd North American

INMAN, D. L. & RUSNACK, G. S . 1956. Changes in sand level on the beach and shelf at La Jolla, California. Technical Memorandum of the United States Army Corps of Engineers, Beach Erosion Board, 82.

MASSEY, B. S. 1970. Mechanics of Fluidr. Van Nostrand Rheinold. MCKEE, B. A., N ~ O U E R , C. A. & DEMASTER, D. J. 1983. Concepts of

sediment deposition and accumulation applied to the continental shelf near the mouth of the Yangtze River. Geology, 11, 631-633.

MCSHEA, D. W. & RAUP, D. M. 1986. Completeness of the geological record. Journal of Geology, 94, 569-574.

NEWELL, N. D. 1972. Stratigraphic gaps and chronostratigraphy. Proceedings of the 24th International Geological Congress, 7, 19&204.

PLOTNICK, R. E. 1986. A fractal model for the distribution of stratigraphic hiatuses. Journal of Geology, 94, 885-890.

REMECK, H.-E. 1960. h e r Zeitliicken in rezenten Flachsee-Sedimenten. Geologkche Rundschau, 49, 149-160.

RETALLACK, G. J. 1984. Completeness of the rock and fossil record; some estimates using fossil soils. Paleobiology, 10, 59-78.

SADLER, P. M. 1981. Sediment accumulation rates and the completeness of stratigraphic sections. Journal of Geology, 89, 569-684. & DINGUS, L. W. 1982. Expected completeness of sedimentary sections:

estimating a time-scale dependent, limiting factor in the resolution of the

fossil record. Proceedings of the 3rd North American Paleontological Convention, 2,461-464.

SCHINDEL, D. E. 1980. Microstratigraphic sampling and the limits of paleontologic resolution. Paleobiology, 6,408-426.

SCHWARZACHER, W. 1975. Sedimentation models and quantitative stratigraphy. Elsevier, Amsterdam.

(Italy). Earth and Planetary Science Letters, 84, 22-26. STRAUSS, D. & SADLER, P. M. 1989. Stochastic models for the completeness

Mathematical Geology, 21, 37-59. of stratigraphic sections. Journal of the International Association for

TIPPER, J. C. 1983. Rates of sedimentation and stratigraphical completeness. Nature, 302, 696-698.

710-715. 1987. Estimating stratigraphic completeness. Journal of Geology, 95,

WEEDON, G. P. 1989. The detection and illustration of regular sedimentary cycles using Walsh Power Spectra and filtering, with examples from the Lias of Switzerland. Journal of the Geological Society, London, 146, 133-144.

WETZEL, A. & AIGNER, T. 1986. Stratigraphic completeness: tiered trace

YALIN, M. S. 1972. Mechanics of Sediment Tramport. Pergamon Press, fossils provide a measuring stick. Geology, 14, 234-237.

Oxford.

- 1987. Astronomically controlled cycles in the lower Tertiary of Gubbio

-

Received 19 October 1988; revised typescript accepted 29 June 1989.

at Tulane University on September 27, 2012http://jgs.lyellcollection.org/Downloaded from