2002 Statistical Appraisal Turbidite JAES 2002 PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Journal of Asian Earth Sciences 21 (2002) 189196 www.elsevier.

com/locate/jseaes

Statistical appraisal of bed thickness patterns in turbidite successions, Andaman Flysch Group, Andaman Islands, India
Partha Pratim Chakraborty*, Basab Mukhopadhyay, Tapan Pal, Tanay Dutta Gupta
Geological Survey of India, Calcutta-16, India Received 20 November 2001; accepted 2 April 2002

Abstract The bed thickness distribution of two turbidite successions in the Oligocene Andaman Flysch Group, India, which differ in terms of their paleogeography and sediment supply, has been assessed statistically. The purpose was to evaluate depositional cyclicity using the nonparametric Waldron test, and also to assess possible environmental control on the bed thickness magnitude distribution. Both sections reveal a positively skewed distribution in bed thickness patterns. The Z value obtained through Waldrons test failed to reject the null hypothesis of randomness and thus discards a cyclic pattern. Juxtaposition of sediments of different proximalities in an elongated forearc fan complex is the possible cause for this absence of cyclicity. The studied sections, however, differ in their bed thickness magnitude distribution. Slope turbidite successions, with sediment supply solely from the outer arc, reveals a scaling behavior while the outer fan succession, with mixing of sediments from different sources, show a negative exponential distribution. This highlights the possible control of the depositional system and sediment supply on the bed thickness magnitude distribution of turbidite successions irrespective of the triggering mechanism. Unequivocal application of the power law distribution in turbidite successions demands reconsideration. q 2002 Published by Elsevier Science Ltd.
Keywords: Turbidite; Andaman Flysch; Depositional cyclicity; Exponential distribution; Power law

1. Introduction: Recognition of cyclicity in turbidite successions has been widely accepted as an important criterion to infer depositional environments of ancient submarine fan deposits (Walker and Mutti, 1973; Martini et al., 1978; Eberli, 1991 and many others). In recent times, identication of cyclicity in lithologic and/or unit thickness variations as well as predictability in their recurrence has resulted in identication of certain deterministic factors for deposition (Goldhammer et al., 1990; Cowan and James, 1996; Diedrich and Wilkinson, 1999). Quantitative and qualitative assessment of the depositional history in ancient deep marine submarine fan sequences through identication of type and scale of cyclicity, however, still remains subjective and heavily biased on the identication of asymmetric cycles, thinning or thickening upward (Waldron, 1987). Detecting these cycles in ancient turbidite sequences partly depends on the
* Corresponding author. Present address: Department of Applied Geology, Indian School of Mines, Dhanbad, India. E-mail address: partha_geology@yahoo.com (P.P. Chakraborty). 1367-9120/02/$ - see front matter q 2002 Published by Elsevier Science Ltd. PII: S 1 3 6 7 - 9 1 2 0 ( 0 2 ) 0 0 0 3 8 - X

choice of beds that dene the base of cycles. Walker (1984) and Hiscott (1981) showed that even a sequence with randomly stacked patterns could generate equally convincing cycles with a positive (thickening-upward) or negative (thinning-upward) trend. In recent times, Chen and Hiscott (1999) criticized the statistical signicance of asymmetric bed thickness cycles as a key criterion for identication of sub-environments in submarine fan systems. Despite being controversial, identication and interpretation of a vertical trend through bed thickness data offers the only methodology to generate reliable estimates and parameters for comparative study of the qualitative dynamics of such deep marine sedimentation settings. Statistical analysis of bed thickness patterns in any stratigraphic succession is principally aimed towards resolving two fundamental aspects of turbidite depositional systems: (i) the degree of randomness of the dataset and, in the case of non-random distribution, quantitative assessment of the dominance of any particular cycle type, positive or negative, and (ii) the physical process of turbidite deposition itselfwhether it shows scale invariance or not. Rothman et al. (1994) emphasized that any turbidite

190

P.P. Chakraborty et al. / Journal of Asian Earth Sciences 21 (2002) 189196

Eocene Oligocene Andaman Flysch Group exposed at Kalipur, North Andaman and Corbyns Cove, South Andaman (Fig. 1), India. These sections differ in lithology (grain size), clast composition and paleocurrent pattern and are inferred to be the product of two different paleogeographic settings viz. slope fan lobe and basin oor outer fan (cf. Chakraborty and Pal, 2001). While the basin oor outer fan deposits reveal a southerly-directed paleocurrent and a substantial amount of quartz (max. up to 28.5%) in the clast population, the slope fan lobe is distinctive in its eastward paleocurrent direction and absence of quartz. This offers an opportunity to verify the applicability and variability of power law statistics in different depositional sub-environments identied independently through observed eld criteria.

2. Geologic setting The Andaman Group of islands form part of an outer arc system with the Indian plate subducting along the Java trench. Bounded between the Mithakhari Group below and Archipelago Group above (Table 1), the 750 m thick (Roy, 1983) siliciclastic Andaman Flysch Group is exposed in detached outcrops over a strike length of 200 km. A longitudinal submarine fan model with a detached slope fan lobe in an accretionary forearc has been proposed (Ray, 1982; Chakraborty and Pal, 2001) for this succession, which shows an onlapping relationship with the underlying Mithakhari Group of trench inner-slope origin (Chakraborty et al., 1999). An active and continuous sediment supply has been interpreted despite a rising or high sea level stand during the development and growth of the fan. Besides being part of two different fan sub-environments, the sections chosen for the present study also differ in terms of their provenance. The accretionary prism is the sole sediment source for the Kalipur section while the Corbyns Cove section represents the mixing of different sources as reected in their different paleocurrent and lithologic characteristics.

Fig. 1. Geological map of the Andaman group of islands; location of studied sections shown.

sequence without any signicant evidence of erosion or amalgamation should t a power law distribution and would reveal a scaling behavior. Unequivocal application of power law statistics to any turbidite bed sequence, regardless of the scale and nature of the depositional system, however, has been contested by Murray et al. (1996) and more recently by Drummond (1999). Control of the depositional environment in terms of turbidite depositional processes was emphasized by Murray et al. (1996) in their study of the Cache Creek section, Northern California, and also by Drummond (1999) through his study of a multi-sourced ramp turbidite succession in the Brallier Formation, Central Appalachian Mountain. This raised a fundamental controversy on whether the scaling observed in many of the turbidite sequences is generic in nature and related to the physical process of turbidite deposition only, or reects the depositional dynamics specic to the depositional environments. Applicability of the power law remains to be tested on submarine fan successions in various tectonic settings. The control of depositional environments on the scaling behavior also needs to be veried. In this paper, we apply Waldrons test and magnitude frequency statistics on two turbidite sections in the Upper
Table 1 Generalized stratigraphy of the Andaman Group of Islands (after Ray, 1982) Stratigraphic Unit (Group) Archipelago Andaman Flysch Mithakhari Ophiolite Unconformity

3. Data collection Special care was taken for demarcation of the top and bottom of the turbidite beds in order to avoid errors in recording the thickness. Besides grain size and sedimentary

Age Pliocene to Miocene Oligocene to Upper Eocene Middle to Lower Eocene Cretaceous

Unconformity/transitional Tectonic/unconformable

P.P. Chakraborty et al. / Journal of Asian Earth Sciences 21 (2002) 189196

191

Fig. 2. Characteristic turbidite section (exposure length 120 m measured along base of the photograph) (a) and representative litholog (b), measured at Coroyns Cove, South Andaman.

structure, positive weathering relief of sandstone and siltstones of turbidity origin relative to the autochthonous inter-turbidite shale units helped us to demarcate and measure the event beds, which were only used for further statistical analysis. Bed thickness measurements were carried out in two sections, 175 km apart, which belong to different fan sub-environments: 199 beds in a mid-fan lobe at Kalipur, North Andaman and 250 beds in an outer fan setting at Corbyns Cove, south Andaman. The minimum thickness recorded in the Kalipur section and Corbyns Cove section is 4 cm and 10 cm respectively. Lithologically, normally graded pebble conglomerate, parallel-sided matrix-supported conglomerate and graded sandstones and laterally continuous shale on the scale of 100s of meters, represent the Kalipur section (Fig. 3). Both normally graded conglomerate and ungraded matrix-supported conglomerate of debris-ow origin, together constituting 15 20% of the total measured section, were measured and included in statistical analysis considering their event origin (cf. Rothman et al., 1994). The Cove section, on the other hand, displays monotonous interlayering of parallel-sided siltstone and shale laterally traceable on a scale of kilometers (Fig. 2). Measurements made in one eld season were veried in two successive eld trips to avoid any bias before carrying out statistical operation.

discussed (Davis, 1986 and many others) of which two methods, viz. Run About Mean (RAM) and Runs up and Down (RUD), can perhaps be regarded as the most powerful statistical tools. These tests not only verify the randomness of the data set in an uninterrupted succession, but also identify the presence of a cyclic pattern (i.e. either thickening upward or thinning upward), if any. Calculation of RAM indicates the dominance of thick or thin beds in a sequence whereas RUD extracts the presence or absence of a local trend and, in turn, the trend in the entire succession. In RUD, the measured succession is coded by 1, 0 or 2 1 by comparing the thickness of the bed with the next bed looking from the bottom to top of a section. If the following bed is thicker than the previous one, then 1 is recorded. Alternatively, if the bed is thinner, then 2 1 is recorded. If both beds are of same thickness, then 0 is recorded. RAM, however, compares the thickness of each bed with the median thickness of the entire succession. When it is found that the bed thickness is greater than the median, then 1 is coded for that bed. If the thickness is equal to the median, then 0 is recorded. If the bed is less than the median, then 2 1 is added. A RAM analysis thus results in the grouping of beds thicker and thinner than the median value. So with these operations, one generates a series of 1, 0 and 2 1 instead of actual bed thickness data. An uninterrupted succession of either 1, 0 or 2 1 can be treated as a run.

5. Bed thickness statistics Waldron (1987) has devised a non-parametric statistical test based on sign differences between consecutive bed thicknesses, to identify the presence of any cyclicity in a measured section, if any, and its statistical signicance. A common observation is that a high noise in the raw data generates a wrong and insignicant result. Elimination of noise from the raw data is therefore a prerequisite to obtain a meaningful result. In the present study, the noise has been reduced by smoothing the raw data by a simple two-bed moving average method (Waldron, 1987; Murray et al., 1996 and others). The Waldron test shows that if a measurement of X number of bed thicknesses generates m number of negative (i.e. 2 1), p number of positive (i.e. 1) transitions and t number of ties (i.e. 0) during the RUD operation where the data is smoothed by a two-bed moving average, then m is normally distributed with the mean

mm N 2 2=2
4. Run analysis A series of continuously measured bed thickness data may be treated mathematically in a Time Series even though the events are not spaced evenly. Taking this into consideration, a number of statistical methods were With a standard deviation q sm N 2=12

where N number of beds (X ) 2 number of ties (t ). A turbidite succession may yield either an upward thinning or upward thickening cycle for which a two-tailed

192

P.P. Chakraborty et al. / Journal of Asian Earth Sciences 21 (2002) 189196

Fig. 3. Bedded matrix-supported conglomerate at the Kalipur section interbedded with rhythmic sandstoneshale (hammer length 27 cm) (a). Representative measured litholog (b) with facies types and process involved are also shown.

test is more appropriate with a null hypothesis that presumes a random distribution of the beds; or alternately, the failure of the hypothesis indicates the presence of an asymmetric cycle. As indicated by Waldron (1987), a continuity correction is needed on m as m is constrained to an integer value whereas its distributions modeled by a continuous function. Thus, if m . mm ; then the corrected value of m0 is equal to m 2 0:5 and when mm , mm then m0 m 0:5:

Now Z m0 2 mm =sm

As Z is normally distributed with mean 0 and standard deviation 1, it lies in between the limits of 1.96 to 2 1.96 a 0:05 for a random succession following the null hypothesis. Any other value of Z supports the alternative hypothesis. The drawback of this test is that it fails when upward thinning and upward thickening cycles are present with equal frequency. Leaving aside such rare instances, in most cases this test successfully determines the major trend in the data. Although the Waldron test was basically designed for RUD analysis, it has been also extended to RAM analysis in order to crosscheck the results.

6. Bed thickness statistics for the cove and kalipur turbidites The 250 bed thickness data (values range from 10 cm to 11 m) of the Cove turbidites show a mean of 1.04 m and standard deviation of 1.22 m. The data indicates a high skewness # 3.892 and kurtosis # 22.92. The 199 bed thickness data of the Kalipur turbidites (values range from 4 cm to 6 m) yield a mean of 60.23 cm and standard deviation of 0.84 m, also exhibiting a high skewness # 3.447 and kurtosis # 15.102. Both data sets on the whole follow a positively skewed distribution. The log-transformation (10 base) of the cove data reduces the skewness value

Fig. 4. Plot of turbidite bed thickness against bed number in (a) Corbyns Cove section and (b) Kalipur section. Note absence of any apparent cyclicity.

P.P. Chakraborty et al. / Journal of Asian Earth Sciences 21 (2002) 189196 Positive transition ( p ) 100, Negative transition ( p ) 96 and Ties (t ) 2 N 199 2 2 197 mm 97:5 and sm 4:0722 As m , mm, then the continuity correction m0 96 0:5 96:5 and Z 20:245563

193

Kalipur turbidite (199 nos. of data)

Positive transition ( p ) 118, Negative transition (m ) 125 and Ties (t ) 6 N 250 2 6 244 mm 121 and sm 4:527692 As m . mm, then the continuity correction m0 2 125 2 0:5 124:5 and Z 0:7730212

Fig. 5. Cumulative frequencythickness (a) and log exceedence probabilitythickness (b) plots of turbidite beds at the Corbyns Cove section. Statistical data (m, a ) indicates slope and intercept respectively.

Kalipur turbidite (199 nos. of data)

Median 39 cm Positive transition ( p ) 99 Negative transition (m ) 96 and Ties (t ) 3 N 199 2 3 196 mm 97 and sm 4:062

As m , mm ; then the continuity correction m0 96 0:5 96:5 and Z 20:12309

Table 2 Comparative results of Cove and Kalipur Turbidite sections under Waldrons test

mm, sm and Z are calculated with the help of the Eqs. (1) (3) (referred in text).

to 0.2109 but generates a x 2 value of 43.03, whereas the same process reduces the skewness value to 0.3648 with a x 2 value of 44.88 in case of Kalipur. The x 2 values are higher than the expected table value and indicate that the bed thickness data of both successions follow a lognormal distribution. To avoid any bias, the plotting of thickness data against the bed number from the base of both measured sections (Fig. 4a and b) were done, which did not reveal any cyclic distribution pattern. Waldrons (1987) test was then applied to both turbidites after the data was smoothed by a two-bed moving average to reduce the noise (Table 2). The Z values in all cases lay within the critical range of 1.96 to 2 1.96 and thus the data of both turbidite successions accept the Null hypothesis. In other words, the null hypothesis of randomness can not be rejected at the 5% level. This failure in rejection of the null hypothesis discards the presence of any asymmetric cycle (thinning or thickening-upward) in either of the two sections studied.

Cove turbidite (250 nos. of data)

RUD analysis

7. Power law or exponential distribution? In order for the power low to hold, the number of beds in a succession thicker than a thickness h must plot as a linear function of h, when plotted in a log log graph (Barton and La Pointe, 1995). If turbidite deposition is essentially guided in a self-organized non-equilibrium system, then the distribution of the thicknesses of turbidite beds that suffered minimal truncation and amalgamation is expected to follow a power law with scale invariance. Thus, within a signicant range of observation, the thickness distribution can be described by a relationship N h ah2B 4

Median 80 cm Positive transition ( p ) 119 Negative transition (m ) 124 and Ties (t ) 6 N 250 2 6 244 mm 121 and sm 4:527692

As m . mm ; then the continuity correction m0 2 124 2 0:5 123:5 and Z 0:55215

Cove turbidite (250 nos. of data)

RAM analysis

Where N is the number of layers thicker than h, B is the scaling exponent and a is a constant. B is an elementary

194

P.P. Chakraborty et al. / Journal of Asian Earth Sciences 21 (2002) 189196

turbidite sections analyzed elsewhere (Rothman et al., 1994; Beattie and Dade, 1996) and explained either through the effect of erosion/amalgamation or as an artifact of the system size. The value of the scaling exponent, calculated through a least-square method on the data and dening a straight line on a log log plot, is found to be 0.943.

8. Discussion The identication of cyclicity with denite trends in ancient submarine fan sequences is considered as important tool for paleogeographic interpretation. Cycles that thin and ne upward are believed to characterize a channel levee succession, while depositional lobes are dened by cycles that thicken and coarsen upward (Shanmugam and Moiola, 1988). Departure from this scheme has also been recorded and interpreted in terms of either retrogressive ow sliding, sea level change (Pickering, 1979; Mutti, 1986) or due to poor development of a point bar and low sinuosity of the channels (Hesse and Dalton, 1995). Conventionally, this paleogeographic modeling based on depositional cyclicity was conned to a siliciclastic radial fan setting with a point source and a dominant shelf supply at low sea level stands. In recent times, active growth of a fan has also been appreciated in transgressive and high sea level stands (Ito, 1998), particularly in deep marine settings. A similar active, tectonically controlled growth in a transgressive sea level stand has also been visualized for the siliciclastic Andaman Flysch fan (Chakraborty and Pal, 2001). A simple fan model is, therefore, not sufcient to justify the sequential trend in fans with (i) multiple sources of sediments, (ii) frequent lobe switching due to changes in local slope conditions and consequent shifting of the feeder channel or, (iii) out of phase interaction between local tectonism and eustatic sea level change with regional variability in the degree of tectonism. The present study failed to identify the presence of any denite cyclical pattern either through a simple bed thickness plot against stratigraphic height or by a nonparametric Waldron test performed on two turbidite sections of the Andaman Flysch fan. This apparent random distribution in bed thickness patterns, as revealed through the sign-of-difference run test, can result from (i) poor estimation due to sectional constraint for identication of any long term steady trend, (ii) aggradational development of the succession, or (iii) presence of a sub equal number of thickening upward and thinning upward cycles in the section. Similar observations from two sections of different paleogeographies rule out the sectional effect in the observed thickness distribution pattern. Aggradational growth may be a possible cause in the fans banked against a fault bounded/ structurally controlled basin margin (Miall, 1989). An aggradational growth pattern can also be important in the detachment of fan lobes from their feeder channels (Shanmugam and Moiola, 1988). However, it is

Fig. 6. Cumulative frequency thickness (a) and their loglog plot (b) in the turbidite succession at Kalipur. Note the scaling behavior and B value less than unity.

statistic that provides a quantitative estimate for turbidite deposition in space and time. In contrast, a Poisson-like process follows a negative exponential distribution and points towards a random behavior of the depositional events. Drummond (1999) summed up the distinction between these two types of distributions through graphical representation using both log normal and log log plots. The distinction between an exponential and a power-law distribution can also be done through an exceedence probability study in magnitude frequency analysis, where the number of beds thicker than any given bed is divided by the total number of beds and that value is plotted against the thickness of the bed (Rothman et al., 1994; Drummond, 1999). The plotting of turbidite bed thickness against the logarithm of their cumulative frequency clearly distinguishes the distribution pattern in the two studied sections: The Corbyns Cove section shows a linear pattern (Fig. 5a), while the Kalipur section reveals a power-law trend (Fig. 6a). The exponential distribution in the Corbyns Cove section is reafrmed through a log exceedence probability plot against turbidite bed thickness (Fig. 5b), which also reveals a linear pattern with a negative slope. The Kalipur section, however, follows the power law distribution for which a number of layers N h of thickness greater than a particular thickness h is plotted against h on log log diagram (Fig. 6b) to see the scaling behavior. Deviations in this plot from a strict linear trend at its upper and lower bounds are similar to those observed in other

P.P. Chakraborty et al. / Journal of Asian Earth Sciences 21 (2002) 189196

195

difcult to visualize aggradation in both the sections, which are spatiality separated and differ both in proximality and sediment input. Bed thickness plots against bed number (Figs. 3 and 4) also rules out the possibility of the presence of both positive and negative cycles. A more plausible explanation may be tectonically induced changing slope conditions, with frequent switching of channels and lobes. Over short time scales, most submarine fans are built by aggradation where channel switching and avulsion causes facies shifts (Chen and Hiscott, 1999). Superposition of products of different proximality might have caused offset in any denite cycle development trend (cf. Einsele, 1991). The thickness frequency statistics of the turbidite beds in the two studied sections, however, show two distinctive distribution patterns indicating that the scaling behavior is not unequivocal. The negative exponential distribution pattern of the Corbyns Cove section is indicative of random behavior of the individual turbidite beds, which might have been incurred through deposition in a Poissonlike stochastic process, where each depositional event is independent of its preceding event. This random behavior is presumably the result of sediment supply from more than one source. Sediment supply in the cove section was mainly from the north along the basin axis with subordinate mixing of materials from the outer arc (Chakraborty and Pal, 2001). Petrographic evidence comes from the presence of quartz in Corbyns Cove sediments, which is absent in the Kalipur section. Drummond (1999) observed a similar exponential distribution in a multi-sourced ramp turbidite succession in the Central Appalachian Basin. The Kalipur section, on the other hand, received sediment exclusively from the outer arc and shows a power law distribution where N scales. The difference in provenance between the two studied sections is further corroborated by their mutually exclusive paleocurrent patterns. One important observation is that the B value for the Kalipur section, which is less than unity. Despite holding the scaling behavior, the observation of a B value more than 1 in both the Izu Bonin fore arc basin and Kingston peak section (Hiscott et al., 1992; Rothman et al., 1994) prompted Rothman and his co worker to visualize the erosion of thin layers or their amalgamation to form thicker layers as possible reasons for their departure from power law statistics. The observed negative B value for the Kalipur section of the Andaman Flysch fan stretches the validity of a scaling limit towards its lower bound. This exercise, however, does not give any idea for the upper bound limit of bed thickness that show scaling behavior. The different distribution patterns in event bed thickness of two turbidite succession belonging to the same tectonic setting, but differing in their provenance and paleogeography, undoubtedly suggests that the depositional system and sediment delivery play crucial roles in modifying their bed thickness frequency patterns. Rothman et al. (1994) envisaged that, except for the system size that modies the scaling factor, seismically induced turbidite successions

should always show the power law distribution. The present study reveals that regardless of whether an earthquake or intrinsic non-linear dynamics of the sedimentary system triggers the turbidite succession, it is the basin geometry and number of sedimentary entry points that have denite control over the nal thickness distribution pattern of the event beds. For a particular point source, sediment delivery can be temporally and volumetrically invariant. However, this may not be valid for a number of sediment entry points, which are likely to be at different geomorphic levels. Qualitative estimations based on B values (i.e sedimentation rate, paleoseismisity) are reliable for turbidite successions in point-source radial fans that follow a power law distribution, but its application is questioned for a multi-source longitudinal fan system. The location of Andaman Island within a subduction environment and the forearc depositional setting of the Andaman Flysch Group suggest a signicant tectonic inuence on the depositional framework as documented by numerous, laterally persistent slump fold layers within the turbidite successions. The inuence of an extrinsic factor such as an earthquake can not be denied, although the depositional system geometry and sediment delivery certainly had control on the bed thickness distribution during the depositional phase.

9. Conclusions Failure of a non-parametric Waldrons test to reject the null hypothesis in two turbidite successions of the Andaman Flysch fan suggests a lack of any depositional cyclicity in either of the two sections studied. Paleogeographic modeling of depositional cyclicity, although proved to be useful for siliciclastic radial fan sequences with a dominant shelf supply, is not adequate for deep marine basins, particularly where the basinal geometry is longitudinal and tectonism plays the principal role in the growth of a fan. The sections, however, differ in their bed thickness distribution pattern. The Corbyns Cove section, which represents the outer fan in a longitudinal fan setting and received sediment from more than one source, shows a negative exponential distribution pattern. Conversely, the slope-fan succession of the Kalipur section, which received sediment only from the outer arc, follows a power-law bed thickness distribution pattern. This observation indicates that despite being part of same tectonic setting and having the same triggering mechanism, the bed thickness distribution patterns in the two turbidite successions may vary depending on the geometry of the depositional system and number of sediment entry points. The present study extends the idea of bed thickness scaling to longitudinal fan settings specically when the sediment delivery is from a point source. On the contrary, a turbidite succession with more than single sediment source is likely to show an exponential distribution where each event is independent of the preceding event.

196

P.P. Chakraborty et al. / Journal of Asian Earth Sciences 21 (2002) 189196 composite sea level changes, cycle stacking patterns and the hierarchy of stratigraphic forcingexamples from Alpine Triassic platform carbonates. Geological Society America Bulletin 102, 535562. Hesse, R., Dalton, E., 1995. Turbidite channel/overbank deposition in a lower Devonian Orogenic shale Basin, Fortin Group of Gaspe/ Peninsula, Northern Appalachians, Canada. Journal of Sedimentary Research B65, 4460. Hiscott, R.N., 1981. Deep sea fan deposits in the Macigno Formation (Middle to Upper Oligocene) of the Gordana valley, Northern Appennines, ItalyDiscussion. Journal Sedimentary Petrology 51, 10151021. Hiscott, R.N., Colella, A., Pezard, P., Lovell, M.A., Malinverna, A., 1992. Sedimentology of deep water volcaniclastics, Oligocene IzuBonin forearc basin, based on formation micro scanner images. Proceedings of the Ocean Drilling Program 126, 7596. Ito, M., 1998. Submarine fan sequences of the lower Kazusa Group, a pliopleistocene fore arc basin in the Boso peninsula, Japan. Sedimentary Geology 122, 69 94. Martini, I.P., Sagri, M., Doveton, J.H., 1978. Lithologic transition and bedthickness periodicities in turbidite successions of the Antola Formation, Northern Appeanines, Italy. Sedimentology 25, 605624. Miall, A.D., 1989. Principles of sedimentary basin analysis. 2nd, 490. Murray, C.J., Lowe, D.R., Graham, S.A., Martinez, P.A., Zeng, J., Carroll, A.R., Cox, T., Hendrix, M., Heubeck, C., Miller, D., Moxon, I.W., Sobel, E., Wendeboung, J., Willams, T., 1996. Statistical Analysis of bed thickness Pattern in a Turbidite Section from the Great Valley Sequence, Cache Creek, North California. Journal of Sedimentary Research 66, 900 908. Mutti, E., 1986. Turbidite systems and their relation to depositional sequences. In: Zuffa, G.G., (Ed.), Provarance in Arenites, Reidel, Dordrecht, pp. 65 93. Pickering, K.T., 1979. Possible retrogressive ow slide deposits from the Kongsfjord Formation: a precambrian submarine fan, Finnmark, North Norway. Sedimentology 26, 295306. Ray, K.K., 1982. A review of the geology of Andaman and Nicobar islands. Miscellaneous Publication of Geological Survey of India 41, 110 125. Rothman, D.H., Grotzinger, J.P., Flemings, P., 1994. scaling in Turbidite deposition. Journal of Sedimentary Research A64, 5967. Roy, T.K., 1983. Geology and hydrocarbon prospects of Andaman Nicobar. Petrol Asia Journal, KDMIPE, ONGC, Dehra Dun, 37 53. Shanmugam, G., Moiola, R.J., 1988. Submarine fans: characteristics, models classication and reservoir potential. Earth Science Review 24, 303 420. Walker, R.G., Mutti, E., 1973. Turbidite facies and facies associations. In: Middleton, G.V., Bouma, A.H. (Eds.), Turbidites and Deep Water Sedimentation, Soc. Econ. Paleontl. Pacc Section Short Course, Anaheim, pp. 119 158. Walker, R.G., 1984. Turbidites and associated coarse clastic deposits. In: Walker, R.G., (Ed.), Facies Models, second ed, Geoscience Canada Reprint Series, vol. 1., pp. 181 188. Waldron, J.W.F., 1987. A statistical test for signicance of thinning- and thickening upward cycles in turbidites. Sedimentary Geology 54, 137 146.

Unequivocal holding of the power law in any turbidite succession should therefore be reconsidered.

Acknowledgments The authors are thankful to Dr S.C. Dasgupta, Director, Andaman Division, GSI for his encouragement in preparation of this paper. They are also indebted to Prof. Pradip K. Bose, Jadavpur University, Kolkata for suggesting improvements in an earlier version of this writing. Reviews from Dr Hugh Sinclair and an anonymous reviewer are thankfully acknowledged.

References
Barton, C.C., La Pointe, P.R., 1995. Fractals in the Earth Sciences, Plenum Press, New York, 265 p. Beattie, P.D., Dade, W.B., 1996. Is scaling in turbidite deposition consistent with forcing by earthquakes? Journal of Sedimentary Research 66, 909 916. Chakraborty, P.P., Pal, T., Dutta Gupta, T., Gupta, K.S., 1999. Facies pattern and depositional motif in an immature Trench-slope basin. Eocene Mithakhari Group, Middle Andaman, India. Journal Geological Society, India 53, 271284. Chakraborty, P.P., Pal, T., 2001. Anatomy of a submarine fan with detached lobe: Upper EoceneOligocene Andaman Flysch Group, Andaman Islands, India. Gondwana Research 4, 477 486. Chen, C., Hiscott, R.N., 1999. Statistical analysis of turbidite cycles in submarine fan successions: Tests for short-term presistence. Journal of Sedimentary Research 69, 486 504. Cowan, C.A., James, N.P., 1996. Autogenic dynamics in carbonate sedimentation: meter scale shallowing upward cycles, Upper Cambrian, Newfounland, Canada. American Journal of Science 296, 11751207. Davis, J.C., 1986. Statistics and Data Analysis in Geology, second ed, Wiley, New York. Diedrich, N.W., Wilkinson, B.H., 1999. Depositional cyclicity in the Lower Devorian Hedelberg Group of New York states. The Journal of Geology 107, 643658. Drummond, C.N., 1999. Bed-thickness structure of multi-sourced ramp turbidite: Devonian Brallier Formation, Central Applachian Basin. Journal of Sedimentary Research 69, 115121. Eberli, G.P., 1991. Calcareous turbidites and their relationship to sea level uctuation and tectonism. In: Einsele, G., Ricken, W., Seilacher, A. (Eds.), Cycles and Events in Stratigraphy, Springer, Berlin, pp. 340359. Einsele, G., 1991. Soumarine mass ow deposits and turbidites. In: Ensele, G., Ricken, W., Seilacher, A. (Eds.), Cycles and Events in Stratigraphy, Springer, Berlin, pp. 313 339. Goldhammer, R.K., Dunn, P.A., Hardie, L.A., 1990. Depositional cycles,

You might also like