From static to dynamic provenance analysis – sedimentary petrology upgraded Eduardo Garzanti PII: DOI: Reference:
S0037-0738(15)00160-8 doi: 10.1016/j.sedgeo.2015.07.010 SEDGEO 4888
To appear in:
Sedimentary Geology
Received date: Revised date: Accepted date:
14 May 2015 9 July 2015 22 July 2015
Please cite this article as: Garzanti, Eduardo, From static to dynamic provenance analysis – sedimentary petrology upgraded, Sedimentary Geology (2015), doi: 10.1016/j.sedgeo.2015.07.010
This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
1
T
ACCEPTED MANUSCRIPT
IP
From static to dynamic provenance analysis –
NU
SC R
sedimentary petrology upgraded
MA
Eduardo Garzanti
AC
CE P
TE
D
Laboratory for Provenance Studies, Department of Earth and Environmental Sciences, Università di Milano-Bicocca, 20126 Milano, Italy
E-mail address:
[email protected]
ACCEPTED MANUSCRIPT
2
ABSTRACT The classical approach to sandstone petrology, established in the golden years of plate tectonics and based on the axiom that "detrital modes of sandstone suites primarily reflect the different tectonic
T
settings of provenance terranes”, has represented a benchmark for decades. The composition of
IP
sand and sandstone, however, simply provides us with a distorted image of the lithological structure
SC R
of source terranes, and gives us little clue whether they are allochthonous or authochthonous, orogenic or anorogenic, young or old. What we may able to see reflected in detrital modes is the nature of source terranes (continental, arc, oceanic) and the tectono-stratigraphic level reached by
NU
erosion in space and time. The proposed new approach to the petrology of sand and sandstone: 1)
MA
starts with a simple classification scheme circulated since the 1960s, which is purely descriptive, objective and free of ill-defined ambiguous terms; 2) focuses on the nature and tectono-stratigraphic level of source terranes. Further steps are essential to upgrade provenance analysis. Acquiring
D
knowledge from modern settings is needed to properly identify and wherever possible correct for
TE
physical and chemical processes introducing environmental and diagenetic bias, and thus address
CE P
nature's complexities with adequate conceptual tools. Equally important is the integration of multiple techniques, ideally including bulk-sediment, multi-mineral and single-mineral methods. Bulk-sediment petrography remains the fundamental approach that allows us to capture the most
AC
precious source of direct provenance information, represented by the mineralogy and texture of rock fragments. Bulk-sediment geochemistry, applicable also to silt and clay carried in suspension, is a superior method to check for hydraulic sorting, chemical weathering and fertility of detrital minerals in different sediment sources. Detrital geochronology, thermochronology and isotope geochemistry reveal the diverse time structures of source rocks, and have become necessary complementary techniques in modern provenance analysis. Inferences on geodynamic processes need independent geological information and come last, but if tackled properly they can lead us much farther than the standard label obtained by using triangular diagrams uncritically as if they were infallible oracles. Keywords: Sandstone petrology; Unroofing trends; Dickinson model; Bulk-sediment methods; Multi-mineral methods; Single-mineral methods.
ACCEPTED MANUSCRIPT
3
T
“Once you have accepted a theory and used it as a tool in your thinking, it is extraordinarily difficult to notice its flaws.” Kahneman, 2011 p.277
SC R
IP
1. Introduction
The origin of sedimentary petrology dates back to the late 19th century, with the invention of thinsection petrography by H.C. Sorby. Provenance studies flourished in the first half of the 20th century, when P.D. Krynine - inspired in Moscow by the ideas of his teacher M.S. Shvetsov -
NU
became the first strong advocate of tectonic control on sandstone composition. A feverish activity in sedimentary petrology followed the first classification schemes proposed in the late 1940s (Krynine,
MA
1948; Pettijohn, 1954), and new classifications continued to proliferate through the 1950s and 1960s (McBride, 1963; Dott, 1964; Folk, 1980). Virtually all of these were projected onto triangular diagrams, which force us to consider three parameters only at a time (quartz, feldspars, rock
D
fragments, or in other cases micas or "clay matrix"). The underlying conceptual schemes were not
TE
based on a thorough investigation of modern settings (Suttner, 1974), and did not benefit yet from the breakthrough of the plate-tectonic revolution. Major progress came with the work of Dickinson
CE P
(1970), who a few years after Gazzi (1966) established operational rules to improve the reproducibility of detrital modes and showed how to connect them with paleogeodynamic scenarios in a seemingly straightforward univocal way (Dickinson and Suczek, 1979). The new paradigm
AC
hinged on the axiom that "detrital modes of sandstone suites primarily reflect the different tectonic settings of provenance terranes” (Dickinson, 1985 p.333). A parallel, somewhat milder statement applied to geochemical composition - echoing Crook (1974) - is Bhatia's (1983 p. 611) “close correlation exists between the geochemical composition of sandstones and tectonic settings of sedimentary basins”. Such claims reflect the enthusiasm of those years, when plate-tectonic theory finally made geologists aware of the fundamental controls of tectonic processes, and seemed able to provide a fresh new explanation for each geological phenomenon. The power of suggestion was strong, and the enchantment persisted through time. Even though other schools refused to adopt the same radical attitude and focused more on compositional modifications during erosion, transport and deposition (e.g., Suttner et al., 1981), sedimentary petrology has remained nailed to that vision for three decades, confirming that “knowledge is in the end based on acknowledgment” (Wittgenstein, 1974 #378). In provenance analysis, as in other fields, we may either trust a simplified model based on what is believed to be essential, or surrender to nature‟s baffling complexities and interpret each case as the
ACCEPTED MANUSCRIPT
4
unique result of multiple competing causes, giving up the quest for a general theory (Paola and Leeder, 2011). To overcome such trade-off, and preserve the benefits that theory offers as a basis for interpretation (Weltje, 2012), we need to acknowledge that the lithological characteristics of
T
parent rocks as inferred from detrital modes of a daughter sandstone cannot be used blindly as a
IP
proxy for plate-tectonic setting. The path leading from a handful of sand to a geodynamic scenario has long been known to be winding and fraught with difficulties (Basu, 1985; Johnsson, 1993). And
SC R
yet, if we make the most of their great potential, then traditional petrographic methods can provide us with a simple and very effective means to identify the nature and tectono-stratigraphy of source terranes, and thus with an unexcelled key to track their erosional evolution through space and time.
NU
Whenever basic tools such as the optical microscope are left aside in favour of advanced efficient machines that produce a great deal of numbers on a very minor and possibly non-representative
MA
fraction of the sediment, the ultimate risk is to find ourselves facing a sea of data without a suitable vessel to sail. The rapid development of technological devices should not make us feel that culture
D
is superfluous, and thus induce us to throw away the traditional keys to understanding.
TE
2. Complexity versus simplicity in provenance analysis French in origin and borrowed from the arts, the term “provenance” refers to the succession of
CE P
passages experienced by a work of art before reaching a museum or a private collection. Provenance analysis, an art by itself, is a thorny business (Zuffa, 1987; Weltje and von Eynatten, 2004). The primary signals imparted by source-rock lithologies undergo diverse physical and
AC
chemical modifications during the sedimentary cycle, both before (“environmental bias”) and after deposition (“diagenetic bias”). Physical and chemical processes produce fundamentally different effects on sediment composition. Whereas the former complicate matters chiefly by distorting the primary signal and can be generally understood and corrected for, the latter may destroy a major and generally unassessable part of the signal, which is thus virtually impossible to restore. Mechanical break-down affects selectively the least durable unconsolidated sedimentary rock fragments (e.g., shale), but not significantly most other framework grains (McBride and Picard, 1987; Garzanti et al., 2015a). Hydrodynamic sorting by size, density and shape controls intrasample and intersample variability, with effects that can be identified and neutralized even in the extreme case of placer lags (Garzanti et al., 2009; Weltje et al., 2015). Chemical weathering, negligible in cold or arid climates (Nesbitt and Young, 1996; Potter et al., 2001; Garzanti et al., 2003), fosters development of thick soil profiles in hot humid climate, with prominent effects (Garzanti et al., 2013a). Corrosion features of detrital minerals offer important
ACCEPTED MANUSCRIPT
5
clues on their relative resistance to weathering (Velbel, 2007; Andò et al., 2012), but tell us the state of what is preserved without helping much to assess what was destroyed. Even quartz and zircon grains suffer from chemical dissolution, and there is no way to recalculate accurately what is gone
T
from what remains.
IP
Most drastic are the effects of post-depositional leaching, because the available time for chemical reactions is much longer, and intrastratal temperatures significantly higher. Many common heavy
SC R
minerals are chemically unstable during burial diagenesis, and are dissolved extensively or even completely in sandstones older than the Plio-Pleistocene. Pyroxene goes first, and next in succession amphibole, epidote, titanite, staurolite and garnet; below 3-4 km only zircon, tourmaline,
NU
rutile, apatite and Cr spinel stand good chances to survive (Morton and Hallsworth, 2007). In most ancient sandstones, heavy-mineral suites consequently represent only the meager durable residue of the much richer and more varied original detrital population. Post-depositional dissolution of main
MA
framework grains may also be extensive, producing “diagenetic quartzarenites” in extreme cases (McBride, 1985). Burial diagenesis is not only an efficient serial killer. During lithification, cement
D
and other authigenic phases may grow in abundance, distorting the orginal composition further,
TE
which represents a particularly serious problem for the interpretation of chemical analyses in the absence of careful petrographic observations. Moreover, chemical effects may be impossible to tell
1967; Dott, 2003).
CE P
from those of recycling, a widespread phenomenon very hard to quantify for sandstone suites (Blatt,
AC
2.1. Starting from the simplicity of a descriptive classification
To bring simplicity in the foreground, we must start anew from the solid basis of accurate description, and get rid of the encrustations of the past, including historical names such as "graywacke" or "arkose" still inexplicably in vogue in geological teaching and literature. The brute term graywacke, referring undecidedly at the same time to colour (gray), texture ("matrix-rich", poorly sorted, angular), composition (lithic-rich but containing quartz and feldspar as well), sedimentary process (typically turbidite) and even age (typically Paleozoic), has generated permanent confusion since the early days of its introduction (Murchison, 1854 p.359; Folk, 1954; Krynine, 1956; Cummins, 1962; Dott, 1964; Dickinson, 1970). Following Folk (1980 p.128), a "graywacke" is nothing else than "a very hard, ugly, dirty, dark rock that you can’t tell much about in field". Arkose is at best an unnecessary ambiguous synonym of feldspar-rich sandstone (Huckenholz, 1963). The attachment to such fantasy names with the idea that "genesis must and does permeate our classification" (Pettijohn, 1948, p.113) retarded analytical investigation by
ACCEPTED MANUSCRIPT
6
involuntarily promoting mythic thinking (sensu Dickinson, 2003). Claiming that genesis should be the basis for understanding is pretending to know beforehand what we want to investigate. Whenever we use "judgments as principles of judgment" we soon find ourselves "going round in a
T
circle" (Wittgenstein, 1974 #124, #191).
IP
The simplest nomenclature, introduced by Crook (1960 p.425) and endorsed by Dickinson (1970 p.697) and Weltje (2006 fig. 2), characterizes sands and sandstones by an adjective reflecting the
SC R
relative abundance of their three main components (quartz, feldspars, lithic fragments). Accessory phyllosilicates or heavy minerals, generally less abundant and markedly affected by hydraulicsorting processes because of their platy shape or high density, are neglected to reduce complexity
NU
and come closer to a transport-invariant measure of sediment composition (Weltje, 2004). The scheme proposed here (Fig. 1) is based on QFL detrital modes obtained by the Gazzi-Dickinson
MA
point-counting method (Ingersoll al., 1984; Zuffa, 1985), modified to record in full detail all encountered rock fragments, which is where the richest and most direct provenance information lies (Suttner and Basu 1985; Garzanti and Vezzoli, 2003). To define aphanitic lithic fragments (L pole)
D
- including all volcanic, ultramafic, metamorphic and sedimentary types and thus also chert,
TE
limestone and dolostone - we use the same 62.5 m cut off (conventional boundary between silt and sand) as Dickinson (1970), whereas Gazzi (1996) proposed originally a 30 m cut off (boundary
CE P
between particles with cohesive and frictional behaviour). The main components, considered only where exceeding 10%QFL, are listed in order of abundance (e.g., in a quartzo-feldspatho-lithic sand L > F > Q > 10%QFL, in a feldspatho-lithic sand L > F > 10%QFL > Q). Because there are as many
AC
rock-fragment types as rocks, useful additional information can be included freely in the definition by a suitable modifier reflecting the most common rock-fragment group (e.g., volcaniclastic, carbonaticlastic, metamorphiclastic, ultramaficlastic; Ingersoll, 1983). A ready objection is that the limited number of variables allowed by graphical displays (three by the equilateral triangle, four by the equilateral tetrahedron projected on the plane) cannot render justice to the richness of detrital suites, including a great variety of rock fragments, feldspars and accessory minerals. To visualize a virtually unlimited number of variables in bi-dimensional space we may recur then to the biplot (Gabriel, 1971). This very efficient statistical/graphical tool helps us not only to discriminate among sample groups, but also to understand the mutual relationships among variables. Parameters other than compositions and of heterogeneous nature can be plotted as well (e.g., grain size in m, distance from the source in km, time in Ma), but the chosen scales of measurement have an impact on the graphical display. The biplot uses the logarithmic transformation, and thus requires that data are all positive (i.e., all zero values in the entry table should be replaced by a suitably small positive number). The principal disadvantage is that each
ACCEPTED MANUSCRIPT
7
addition of a variable or of a sample modifies the biplot. Provenance fields of reference, therefore, cannot be defined as done commonly with Cartesian or triangular diagrams.
IP
T
2.2. Unraveling complexity by exploring modern Earth
Our chance of making correct provenance diagnoses becomes small when geological age blurs our
SC R
landmarks, and the complexity of natural systems confounds us in a labyrinth of possibilities. The only hope to carry out the task successfully is to learn as much as we can from modern cases, where we do not have to worry about the superposed effects of diagenesis and full geological and
NU
geomorphological information on source terranes and sediment-routing systems is available, thus providing suitable conditions in which each control on sediment composition can be identified and
MA
individually quantified (Ingersoll, 1990; Le Pera and Critelli, 1997). The study of modern settings helps us to perceive the complexities of source-to-sink dispersal paths, and the difficulties involved in the reconstruction of ancient landscapes (Allen, 2008; Hinderer, 2012). Only by comparison with
D
the modern can we understand and correct for hydraulic-sorting effects, evaluate the importance of
TE
weathering and recycling, realize the uncertainties and potential pitfalls nested in our thinking, and
CE P
ultimately avoid moving in circle in the attempt to confirm our prejudices. 3. The integrated approach to provenance diagnosis
The best strategy to tackle complexity is to combine a diversified set of analytical methods and
AC
provenance tracers (Najman, 2006). Different techniques provide distinct points of view, from which disparate details of the general picture can be revealed. Only by the painstaking careful integration of such diverse complementary pieces of information can we hope to get a glimpse of the entire landscape. 3.1. Implicitly chosen grain-size windows
Terrigenous sediments range from clay to boulders, spanning many orders of magnitude (i.e., from few microns to several meters). There is hardly a single method applicable to such a huge spectrum of grain sizes. While choosing a certain analytical approach we must remain fully aware that we have consequently restricted our focus to the limited - sometimes very limited - part of the grainsize spectrum that can be investigated by the chosen technique. The harder we work on that size window, the easier it becomes to forget that we are in the meantime disregarding the rest.
ACCEPTED MANUSCRIPT
8
Most techniques work well on sand only. Gravel must be generally tackled in the field, with limited means. Muds and mudstones, representing the largest part of the sedimentary record (Blatt, 1985), are hardly dealt with by optical methods or single-grain techniques, and geochemistry, X-ray diffraction or
T
Raman spectroscopy are generally used to monitor suspended load, which makes up the bulk of fluvial
IP
sediment transport (Bangs-Rooney and Basu, 1994; Andò et al., 2011; Bouchez et al., 2011). Geochemistry is the only flexible technique that can be equally employed on clay, silt, sand and
SC R
granules, and therefore on most components of the sediment flux (von Eynatten et al., 2012). Bulk petrography has a more limited range, being well suited to sand-sized sediments only. Information quality and analytical precision decrease quickly from very fine sand to very coarse silt, and little can
NU
be done on cohesive mud. Very-coarse sand and fine gravel are suited only for the detailed analysis of rock fragments, because thin sections are too small to contain a representative sample of such coarse
MA
clasts. Pebbles, cobbles and boulders have to be studied one by one, which may provide precious information on the lithologies that yield specific detrital minerals, on the orogenic versus anorogenic geochemical affinity of volcanic parent rocks, or on the age and exhumation histories of plutonic and
D
metamorphic source terranes (Bluck et al., 2006; Dunkl et al., 2009; Spalla et al., 2009). Where
2013b; Limonta et al., 2014).
TE
sandstone clasts occur, within-clast point-counting is essential in the study of recycling (Garzanti et al.,
CE P
Heavy-mineral analyses can be performed down to medium silt size under the optical microscope, and almost in the entire silt range with the help of Raman spectroscopy (Andò and Garzanti, 2014). Singlemineral techniques are applied to a more limited part of the grain-size spectrum (typically very coarse silt to lower medium sand; Lawrence et al., 2011; von Eynatten and Dunkl, 2012). Focusing on one
AC
single detrital-mineral population (e.g., zircons), representing a minimal part of a minor part of the total sediment flux (e.g., sand bedload), brings on the risk of ignoring the bulk of what lies out there in nature (Weltje and von Eynatten, 2012).
3.2. Bulk-sediment methods
Because of the limitations implicit in any technique, in most provenance studies we end up to analyze sets of sediment samples that may be scarcely representative of the entire spectrum of grain sizes contained in the total sediment flux. To avoid making things worse by analyzing sub-samples not even representative of the sample from which they were extracted, bulk-sample approaches should be the starting point of any provenance investigation. Taking into account the largest possible fraction of total transported sediment is a pre-requisite to avoid the pitfalls of source-rock
ACCEPTED MANUSCRIPT
9
fertility, and to calculate sufficiently accurate sediment budgets from which average erosion rates can be derived. Traditional petrography is by far the most straightforward, cheap and efficient means to determine
T
the mineralogical and textural parameters characterizing the sand-sized sediment under
IP
examination. Rich information on parent rocks is revealed directly by the detailed study of rock fragments, which provide us with a firm basis for provenance diagnosis. In particular, the rigorous
SC R
definition and systematic determination of metamorphic rank for each counted metapelite, metapsammite and metabasite rock fragment allows us to estimate the average metamorphic grade of parent rocks (MI index of Garzanti and Vezzoli, 2003), and hence the crustal level reached by
NU
erosion in source areas (Garzanti et al., 2006, 2010b). Although the recent tendency is to rely more and more on sophisticated instruments for the collection of numerical datasheets, the use of
MA
valuable data-acquisition machinery does not automatically lead to valuable science, and basic examination of thin sections under the microscope remains very necessary. One microscope and thin sections are often enough to look at a sand-sized sediment or sedimentary
D
rock and unveil many of its secrets, even though the famed Kryninism "give me one thin section
TE
and I'll give you the story of the Appalachian geosyncline" (Bates and Griffiths, 1971) is received as an obvious overstatement. There are quite a few important provenance features that petrography
CE P
cannot capture. Without the help of detrital geochronology nothing is seen of the time structures of source terranes, and we cannot guess whether they are young or old, allochthonous or autochtonous, and thus orogenic or anorogenic. There is no ready way to discriminate detritus derived from sections of continental crust exposed along a rifted margin or incorporated tectonically in the
AC
external belt of an orogen, from neometamorphic or paleometamorphic sources within the same orogen, from a neometamorphic axial belt or an old cratonic crust, from recent or old ophiolitic sutures, from active neovolcanic and inactive paleovolcanic arcs or anorogenic volcanic fields. Geodynamic setting, therefore, cannot be inferred directly from detrital modes of sand and sandstone. Bulk-sediment geochemistry represents a suitable complementary approach, and the wide spectrum of chemical elements and their different behavior provide invaluable information in sediment-generation studies. The abundance of ultradense species (e.g., zircon) can be calculated approximately from the concentration of elements hosted principally in that mineral (e.g., Zr, Hf), which makes geochemistry a most efficient tool to assess the fertility of different sediment sources (Dickinson, 2008) as well as hydraulic-sorting effects leading to the anomalous concentration of such minerals in different samples or size classes (Garzanti et al., 2010a, 2011). Most chemical elements, however, are hosted in significant proportions in diverse detrital minerals, which blurs the geochemical fingerprint of specific
ACCEPTED MANUSCRIPT
10
sources while detritus gets progressively homogenized during downstream transport in higher-order rivers to the sea (Ingersoll, 1990; Garzanti et al., 2014). Moreover, those primary signals may be effaced by orders-of-magnitude stronger hydrodynamic effects, and bulk-sediment geochemistry is
T
thus generally revealed as a rather blunt tool for provenance diagnosis. A happy exception is
IP
represented by elements hosted preferentially in mafic and ultramafic rocks (i.e., Cr, Ni and to a lesser extent Mg; von Eynatten et al., 2003), which carry a signal commonly strong enough to survive
SC R
downcurrent homogenization and environmental bias (Amorosi, 2012; Garzanti et al., 2012). Sediment geochemistry is also widely used to infer the extent of chemical weathering (Nesbitt and Young, 1982; Price and Velbel, 2003; Borges et al., 2008; Shao et al., 2012), although weathering, hydrodynamic,
NU
grain size and provenance effects must be detangled carefully to avoid misinterpretations (Bloemsma et al., 2012; Garzanti and Resentini, 2015; von Eynatten et al., 2015). The extensive dissolution of
MA
unstable minerals and massive precipitation of cements and authigenic material during burial diagenesis may considerably reduce the usefulness of bulk-sediment geochemistry in the provenance study of ancient sandstones. 87
Sr/86Sr,
143
Nd/144Nd) provide important information on crustal evolution
D
Isotope ratios (e.g.,
TE
(Goldstein and Jacobsen, 1988) and are generally less sensitive to environmental and diagenetic bias. The relative sediment contribution from two sources with very distinct isotopic fingerprints
CE P
can be assessed accurately, but where detritus is derived from multiple tectonic domains with overlapping signatures, as commonly is the case in orogenic settings or large river systems, then single isotopic ratios give equivocal responses and can serve us at best to check provenance estimates made independently with other methods (Clift et al., 2002; Padoan et al., 2011).
AC
Moreover, we must decipher the relative role played by the detrital minerals that control the isotopic budget of bedload and suspended load in fluvial systems (Garçon et al., 2014).
3.3. Multi-mineral methods
A complete panorama of the great potential of heavy-mineral studies is provided in the monumental book by Mange and Wright (2007) and in the detailed updated review by Morton (2012). The investigation of multi-mineral suites reveals crucial provenance information for paleotectonic reconstructions (e.g., Dewey, 2005), especially if coupled with petrographic observations. Two fundamental aspects for the correct description and interpretation of heavy-mineral assemblages, whether they are analyzed optically or by more advanced techniques (e.g., microprobe, QemScan), are however overlooked frequently.
ACCEPTED MANUSCRIPT
11
The first aspect concerns the practical need to select a specific grain-size window for analysis wherever the presence of detrital grains with great size differences makes laboratory procedures troublesome (Mange and Maurer, 1992). Under the wrong implicit assumption that grain-size
T
classes represent transport-invariant subpopulations, and with the illusion that narrowing the size-
IP
window increases consistency whereas in fact it is prone to maximize bias, several authors recommended to analyze a single class not more than 1 or even only 0.5wide (125-250 μm,
SC R
Carver, 1971; 63-125 μm, Morton, 1985; 90-125 μm, Bateman and Catt, 2007). Because highdensity minerals settle at the same velocity of - and hence are deposited together with - much coarser low-density or platy minerals, the former concentrate in the fine tail of any sorted sediment,
NU
the latter in the coarse tail (Rubey, 1933; Garzanti et al., 2008). The different grain-size classes of a sorted sediment, therefore, have notably different composition, and bulk-sample or multiple-
MA
window analyses represent the only correct options to estimate the percentages of detrital minerals accurately. Provenance diagnoses or stratigraphic correlations cannot be made any more accurate at either local or regional scale by considering a homogeneously narrow size range for a texturally
D
inhomogeneous suite of terrigenous rocks. To minimize analytical bias, well-sorted sands should be
TE
analyzed in bulk, whereas the widest possible size-window centered about the mean should be chosen for poorly-sorted sediments (Garzanti et al., 2009). Point-counting techniques are highly
CE P
recommended for heavy-mineral analyses carried out on bulk-samples or wide grain-size windows. This is because the discrepancy between the real volume percentages and the number percentages as determined by grain counting increases with the increasing width of the grain-size window analyzed, and volume percentages are systematically overestimated for denser minerals that are
AC
smaller than settling-equivalent lower-density minerals (Galehouse, 1971). Image analysis represents a useful complementary technique to tackle such operational problems. The second aspect concerns the need to estimate not only the relative abundance of diverse heavy minerals but also their absolute abundance (i.e., their concentration in the bulk sample; Garzanti and Andò, 2007a). The concentration (and not just the spectrum) of heavy minerals in a sedimentary deposit depends on the composition of parent rocks, and increases by more than an order of magnitude during progressive unroofing of denser rocks found at deeper-seated crustal levels (Garzanti et al., 2006). Equally drastic modifications of their concentration (as well as of their spectrum) may occur by selective entrainment of low-density grains in the depositional environment (Garzanti et al., 2010a), or by selective leaching of unstable species during diagenesis (Gazzi, 1965; Andò et al., 2012). The concentration of heavy minerals (as well as their spectrum) is therefore per sè crucial in provenance interpretations and in the correct assessment of recycling, hydraulic and diagenetic processes. The distortive fertility effect related to the different potential of different rock
ACCEPTED MANUSCRIPT
12
types to generate heavy minerals must always be taken into full account in the interpretation of heavy-mineral suites, which tend to document aberrantly a limited number of sources (e.g., mafic igneous and metamorphic rocks), whereas several others are barely recorded (limestone, chert,
T
shale, granite). In the absence of significant weathering and diagenesis, mafic rocks may thus
IP
impose their mark on the heavy-mineral spectrum even where their outcrops are sparse (figure 1 in
SC R
Garzanti and Andò, 2007a). 3.4. Single-mineral methods
An excellent comprehensive synthesis of the advanced techniques and target minerals used in
NU
modern provenance analysis is provided in von Eynatten and Dunkl (2012). Minerochemical methods can be applied to any detrital species displaying significant compositional variability,
MA
whereas geochronological and thermochronological methods are limited to the narrower range of suitable minerals containing unstable isotopes. Fractionation by physical or chemical processes is less fastidious in the case of single-mineral approaches, because size, shape, density and chemical
D
durability vary less within a single-mineral population than within the entire detrital population.
TE
After the advent of detrital geochronology and thermochronology we can now investigate the diverse time structures of source terranes (Vermeesch et al., 2009), a powerful complement to the
CE P
traditional petrographic or geochemical approaches that provide information on the lithological structure of parent rocks only. Zircon, widespread in recycled sands and diagenized sandstones because of its durability, is the most commonly targeted mineral. U-Pb age spectra of detrital
AC
zircons reflect the crystallization ages of exposed magmatic and metamorphic rocks, whereas their fission-track age and length distributions - if not reset after deposition - are related to exhumation mode and timing in parent-rock domains. Zircon-rich felsic rocks, however, will show up much better than zircon-poor ones even though their outcrop area or erosion rate are limited, and other source rocks will not show up at all, including basalt, serpentinite, carbonate or chert. Because of zircon durability, age spectra may remain unchanged through successive recycling episodes and consequently homogeneous in time and space, ceasing to be useful provenance tracers (Garzanti et al., 2013c). And even zircon populations may be fractionated by hydrodynamic processes (Lawrence et al., 2011). But the most insidious peril of single-mineral approaches is that the researcher may end up to believe that “what he sees is all there is” (Kahneman, 2011 p.85). Provenance inferences obtained from a single mineral cannot be extrapolated to the bulk sediment unless its fertility in all potential sources is known accurately (Moecher and Samson, 2006; Malusà et al., 2015), which is hardly ever the case. Zircon has exceptionally useful properties, but its average content in sediments is only 2
ACCEPTED MANUSCRIPT
13
grains out of 10,000 (roughly corresponding to 200 ppm of Zr in the upper continental crust; Taylor and McLennan, 1995). By focusing on zircon exclusively we shall miss information from the remaining 99.98% of the sample. Exploring a very minor portion of the explorable may not turn out
T
to be the clever strategy. While a powerful beam of light illuminates a small detail, the rest of the
IP
picture remains in darkness.
SC R
4. The plate-tectonic paradigm
The plate-tectonic revolution of the 1960s and 1970s offered a new elegant conceptual framework
NU
and opened up novel ways to explore the history of continents and oceans. Past geodynamic scenarios could now be unveiled by trace-element signatures of igneous rocks (Pearce and Cann,
MA
1973; Winchester and Floyd, 1977), and the intuition that the same could be done with the compositional signature of terrigenous sediments inspired the outstanding contribution of the Dickinson school in the 1970s and 1980s, as well as various geochemical studies (Bhatia, 1983;
D
Roser and Korsch, 1986). The hope that the mark of ancient plate-tectonic settings could be testified
TE
univocally by, and thus inferred unequivocally from detrital modes of sandstones gave rise to great scientific interest and a boom of sandstone-petrology studies in successive years. By emphasizing
CE P
plate-tectonic control, however, the new model pushed to the background the compositional modifications induced by physical and chemical processes during the sedimentary cycle, and all details of sediment generation. Moreover, the geodynamic framework of reference was conceptually simplified to the point that disillusion inevitably followed, and in successive decades
AC
petrographic methods were progressively abandoned in favor of more sophisticated approaches in the quest of those magical solutions that traditional techniques seemed unable to provide. Criticism of the model came early, but was largely limited to specific exceptions (Mack, 1984) rather than on more substantial flaws, such as the oversimplification of orogenic domains and the disregard of rift-related settings, anorogenic volcanism and ophiolitic sources. Carbonate rock fragments, widespread in all but the wettest monsoonal and equatorial climates (Zuffa, 1985, 1987), were largely neglected. Chert was considered typical of Subduction Complex provenance, which is contradicted by studies of modern depositional systems (Garzanti et al., 2002a, 2013b). But above all, sandstone petrography alone cannot tell us the age of igneous or metamorphic source rocks, or whether they are allochthonous or autochtonous. Orogenic and anorogenic provenances, therefore, can hardly be discriminated (Fig. 2), and geodynamic setting cannot be univocally inferred from detrital modes of sandstones.
ACCEPTED MANUSCRIPT
14
Another major problem with the Dickinson model is that it was built mostly on ancient sandstone suites, more than 80% and 90% of clastic units considered respectively in Dickinson and Suczek (1979) and Dickinson et al. (1983) being older than the Neogene. Because ancient geodynamic
T
settings are assumed rather than known, the model is partly based on circular reasoning. The
IP
rigorous analysis by Weltje (2006), following that of Molinaroli et al. (1991), shows that the identified Continental Block, Magmatic Arc and Recycled Orogen provenances can be
SC R
distinguished by Dickinson‟s plots with a success rate of only 64% to 78%. These considerations indicate clearly that, in spite of its fame and faithful application in so many provenance studies, the model cannot be used blindly as an oracle (and even less as a classification).
NU
The mistake, however, does not reside in the model but rather in its uncritical use. Models are not meant to be infallible, but just conceptual tools built because they are easier to handle than nature
MA
itself (Borges, 1960). As for spherical cows (Paola and Leeder, 2011), we should appreciate the theoretical simplification rather than taking them as gospel truth. Dickinson himself (1985 p. 351; 1988 p.9) advised to use ternary diagrams as graphical devices to illustrate mixing of detritus from
D
diverse end-member sources, rather than rigidly as discrimination oracles of universal validity. Such
TE
a flexible creative approach allows us to describe sediment mixing in space and time (Dickinson, 1985 p.352-354), thus moving forward from the standard choice of a static paleogeographic
CE P
scenario to the methodical definition of unroofing trends reflecting the dynamic evolution of diverse crustal sources (Fig. 3).
AC
4.1. From static to dynamic sedimentary petrology The seminal article by Dickinson and Suczek (1979) opened the door for plate-tectonics to enter the field of sedimentology, and plate-tectonic control on sediment composition has remained the dominant paradigm for the subsequent decades. The model works excellently for magmatic arcs, the geodynamic setting most thoroughly investigated in successive years (Marsaglia and Ingersoll, 1992 and references therein). Magmatic arcs are relatively easy to model as commonly isolated simple systems. The classical unroofing trend from volcaniclastic to plutoniclastic detritus associated with progressive deepening of erosion into the arc massif, defined originally in the Great Valley forearc sequence (Dickinson and Rich, 1972; Ingersoll, 1983), confirmed subsequently its validity even in more complex orogenic settings (Garzanti et al., 1996). Such an application of triangular diagrams to predict changing detrital modes during unroofing of the source terrane (e.g., figure 3 in Dickinson, 1985) indicated the way to tackle other anorogenic and orogenic provenances with the same dynamic approach (Garzanti et al., 2001, 2002b, 2004, 2010b). Considering the evolution of source areas in time provides us with a richer perspective in provenance analysis, leading us to
ACCEPTED MANUSCRIPT
15
examine the stratigraphic record carefully and to look for the trace of eroding source terranes within changing paleogeographic scenarios (e.g., table 2 in Garzanti et al., 2012). To make a step forward, however, some of the tenets on which the classic paradigm is based need to
T
be abandoned. First of all, the illusion that sediment composition is controlled strictly by plate-
IP
tectonic setting, whereas in reality sediment composition simply provides us with a distorted image of the lithological structure of source terranes. Second, the illusion that sediments generated in
they simply do not (Fig. 2; Fig. 3).
NU
5. Tectono-stratigraphic levels and unroofing trends
SC R
different settings should plot diligently in separate places within a QFL diagram, whereas in reality
The idea that sediment composition is primarily controlled by the tectono-stratigraphic level
MA
undergoing erosion in the source area was first expressed clearly by Krynine (1948), who envisaged the continental crust as consisting of three layers: sediments on top, metamorphic rocks and veins
D
next, and plutonic igneous rocks at depth. With increasing amounts of tectonic activity (and/or time) these successively deeper layers are brought to the surface and act as a source for sediments,
TE
yielding in succession "quartzite", lithic-rich "graywacke" and feldspar-rich "arkose" (Folk, 1980 p.108). A similar partition is found in Continental Block provenance, subdivided by Dickinson and
CE P
Suczek (1979 p.2175) into Craton Interior subprovenance characterized by quartzose sand with low P/F ratio, and Uplifted Basement subprovenance characterized by quartzo-feldspathic sand. The occurrence of more lithic sand wherever erosion fails to completely remove supracrustal
AC
sedimentary or metamorphic rocks overlying crystalline basement was acknowledged, and Transitional Continental subprovenance formalized later on (Dickinson et al., 1983). These three subdivisions, broadly equivalent to Krynine's and confirmed by studies of modern rift settings, define an unroofing trend (Garzanti et al., 2001). To emphasize this aspect, the corresponding labels "Undissected", "Transitional" and "Dissected" can be used also for Continental Block provenance, and not for Magmatic Arc provenance only. Not only arc crust (Dickinson and Suczek, 1979) or continental crust (Garzanti et al., 2006), but even ophiolitic allochthons including oceanic crust and underlying mantle can be envisaged as a progressively eroding multilayer source of sediment (Ophiolite provenance; Garzanti et al., 2002b). The expected petrofacies successions associated with unroofing of deeper tectono-stratigraphic levels in these three distinct cases define the building blocks of the dynamic provenance model presented here in its essence. How to deal with the complexities caused by mixing of sediments with different provenances in anorogenic and orogenic settings will be briefly discussed next.
ACCEPTED MANUSCRIPT
16
5.1. Unroofing of arc crust
As envisaged in the simplest way, arc crust consists of andesitic volcanic covers made of vitric
T
groundmass, plagioclase and pyroxenes, overlying tonalitic/granodioritic batholiths made of quartz,
IP
plagioclase, K-feldspar, biotite and hornblende. Ideal unroofing trends are thus characterized by a progressive increase in quartz, K-feldspar and hornblende at the expense of volcanic lithics and
SC R
pyroxenes, with consequent transition from feldspatho-lithic volcaniclastic sand rich in pyroxenes (Undissected Magmatic Arc provenance) to quartzo-feldspathic plutoniclastic sand rich in amphibole (Dissected Magmatic Arc provenance; Marsaglia and Ingersoll, 1992; Garzanti and
NU
Andò, 2007b).
MA
5.2. Unroofing of continental crust
Continental crust has a heterogeneous polymetamorhpic structure inherited from a series of
D
previous orogenic cycles. It can be idealized as a tectono-stratigraphic multilayer composed, from
TE
top to bottom, of unmetamorphosed cover strata, upper-crustal anchimetamorphic to greenschistfacies metasediments, middle-crustal amphibolite-facies gneisses and granitoids, and lower-crustal
CE P
granulites and mafic intrusions. Metamorphic grade, rock density, abundance of plutonic protoliths, and ratio of mafic to felsic products all tend to increase from shallower to deeper tectonostratigraphic levels (Handy 1990).
Sands derived from continental blocks have been classically thought of as consisting chiefly of
AC
quartz and feldspar, and thus fundamentally distinguished according to their quartz/feldspar ratio, that increases with weathering or recycling and decreases with active tectonic uplift (Folk, 1980 p.130-133; Dickinson, 1985). Sands released from supracrustal rocks, however, may range from quartzose to lithic sedimentaclastic (Fig. 2), including a variety of rock-fragment types and very poor heavy-mineral suites dominated by zircon, tourmaline and rutile (Undissected Continental Block provenance). Mid-crustal crystalline basements shed instead feldspatho-quartzose to quartzofeldspathic sand with rich hornblende-dominated suites (Dissected Continental Block provenance). The Metamorphic Index MI increases steadily when and where erosion cuts deeper into the crustal multilayer (Garzanti et al., 2006). 5.3. Unroofing of oceanic lithosphere Tectonically accreted or obducted sections of oceanic lithosphere that escaped subduction and orogenic metamorphism consist of thin abyssal siliceous oozes mantling pillow basalts made of
ACCEPTED MANUSCRIPT
17
vitric groundmass, plagioclase laths and pyroxene, overlying in turn diabase dykes yielding epidote and actinolite grown during oceanic metamorphism. The underlying gabbros or gabbro-norites, with small plagiogranite bodies at the top and ultramafic cumulates at the base, are mostly made of calcic
T
plagioclase, pyroxenes, hornblende and olivine. Commonly harzburgitic mantle peridotites
IP
represent the base of the ophiolitic multilayer. During unroofing, feldspatho-lithic basalticlastic sand rich in lathwork volcanic grains and pyroxene (Undissected Ophiolite provenance) is ideally
SC R
replaced by lithic ultramaficlastic sand dominated by lizardite-serpentinite rock fragments and rich in enstatite, olivine and minor Cr-spinel (Dissected Ophiolite provenance; Garzanti et al., 2002b). Orthopyroxene-phyric
boninite grains may be common in detritus from
undissected
MA
5.4. Mixed provenance in anorogenic settings
NU
suprasubduction-zone ophiolites (Fig. 3; Garzanti et al. 2000).
Lithologic assemblages exposed at divergent plate margins include continental basements and cover
D
rocks stripped down to various levels along the rift shoulders (Continental Block provenance), as well as rift-related volcanic rocks shedding feldspatho-lithic volcaniclastic sand rich in
TE
clinopyroxene (Anorogenic Volcanic provenance; Fig. 2). Where pull-apart basins are tectonically inverted during late rifting stages, recycling of syn-rift clastic successions may produce quartz-rich
CE P
sands with poor heavy-mineral suites (Recycled Clastic provenance; Garzanti et al., 2014). Sediments generated in anorogenic settings, such as those found in diverse tracts of the huge Nile
al., 2015b).
AC
catchment, can thus be modeled as mixtures of these three end-member provenances (Garzanti et
5.5. Mixed provenance in orogenic settings Modeling orogenic detritus poses difficult problems, solved in the Dickinson model by the loose definition of three orogenic settings and corresponding subprovenances (subduction complex, collision orogen/suture belt, foreland uplift/fold-thrust belt) lumped under the generic label of Recycled Orogen provenance (Dickinson and Suczek, 1979). Because of unclear operational criteria, such subcategories were seldom used subsequently. A more articulate solution is based on two successive logical steps (Garzanti et al., 2007). First, a limited number of orogen archetypes are identified (Himalayan-type collision orogens, Andean-type cordilleras, Oman-type obduction orogens, Apennine-type thin-skinned belts, Indo-Burman-type subduction complexes). These are not just variants within a single orogenic spectrum, but represent different geological objects generated by subduction processes with different geometry and involving plates of different nature
ACCEPTED MANUSCRIPT
18
(continental or oceanic) in the footwall or hangingwall. Consequently, they are made of different materials and shed sediments with different composition. Second, orogens are seen as resulting from the juxtaposition and superposition of a limited number of tectonic domains arranged in
T
subparallel linear belts that jointly contribute mixed detritus to the associated sedimentary basins
IP
(Dickinson and Suczek 1979 p.2176). Five types of such elongated geological domains can be identified as the primary building blocks of all composite orogens (magmatic arcs, obducted or
SC R
accreted ophiolites, neometamorphic axial belts, accreted remnants of rifted continental margins, accreted clastic wedges), each producing predictable detrital modes, heavy-mineral assemblages and unroofing trends (Fig. 3). Such five primary provenances (Magmatic Arc, Ophiolite, Axial Belt,
NU
Continental Block, Recycled Clastic) can be recombined to describe the full complexities of mixed
5.6. The upgraded provenance model
MA
detrital signatures produced by erosion of different types of orogenic prisms.
D
The provenance model presented here considers several essential aspects neglected in the Dickinson model (i.e., rift-related settings, anorogenic volcanism, ophiolitic allochthons, recycling), but
TE
includes only six (three more) primary provenance types (Continental Block, Magmatic Arc, Ophiolite, Axial Belt, Anorogenic Volcanic, Recycled Clastic). Continental Block and Recycled
CE P
Clastic provenances are common to both anorogenic and orogenic settings, because detrital modes cannot distinguish between authochtonous and allochthonous continental crust, and because recycling may occur along both divergent and convergent plate margins. Recycled Orogen
AC
provenance is dropped in favor of Axial Belt provenance, which has a more restricted definition (Garzanti et al., 2010b).
The dynamic upgraded model allows for progressive changes of sediment mineralogy in space and time during the erosional evolution of the same source area. The trends of unroofing defined for different primary provenances show wide overlap on the QFL diagram (Fig. 3), and several converge towards quartzo-feldspathic and hornblende-rich "ideal arkose" resulting from erosion of mid-crustal granitoid rocks similarly exposed in dissected continental blocks, dissected arc massifs or dissected orogenic belts (Dickinson, 1985; Carrapa and Di Giulio, 2001). The QFL plot thus cannot be used to infer provenance directly, but remains valid to outline mixing of detritus from diverse end-member sources (Dickinson, 1988), or as a basic graphic tool for classification (Fig. 1). The most robust key to provenance diagnosis lies in the rich mineralogical and textural spectrum displayed by rock fragments, as in the equally varied heavy-mineral suites where not depleted drastically by diagenetic dissolution.
ACCEPTED MANUSCRIPT
19
6. Concluding remark Detrital modes of sandstones cannot tell us the plate-tectonic setting in which they were produced, because petrography alone cannot discriminate allochthonous versus authochthonous, orogenic
T
versus anorogenic, or young versus old parent rocks. If we pay detailed attention to rock fragment
IP
types and to both relative and absolute heavy-mineral concentrations, then sandstone composition
SC R
can reveal the tectono-stratigraphic level reached by erosion in continental, arc and oceanic sourcerock domains. If we set ourselves free from obsolete or simplistic views and refrain from mythic thinking (Dickinson, 2003), then plate-tectonic processes and their evolution in space and time can
NU
be eventually unveiled by using an integrated set of bulk-sediment, multi-mineral and singlemineral techniques, coupled with the indispensable complementary geological information.
MA
ACKNOWLEDGMENTS
Hilmar von Eynatten and Istvan Dunkl kindly invited me to present these thoughts in the form of a
D
talk at the 2°WSGS held in Goettingen in June 2014. The ideas exposed herein stem from fruitful
TE
continuing collaboration with many friends around the world including Yani Najman, Pieter Vermeesch and Xiumian Hu, and enjoyable strenuous work carried out through the years with
CE P
Sergio Andò, Giovanni Vezzoli and our beloved undergraduate and postgraduate students at Milano-Bicocca. Careful acute reviews by Abhijit Basu and Gert Weltje, and precious comments by
AC
Hilmar von Eynatten and Alberto Resentini, are gratefully acknowledged.
ACCEPTED MANUSCRIPT
20
REFERENCES Allen, P.A., 2008. From landscapes into geological history. Nature 451, 274-276.
T
Amorosi, A., 2012. Chromium and nickel as indicators of source-to-sink sediment transfer in a Holocene alluvial and coastal system (Po Plain, Italy). Sedimentary Geology 280, 260-269.
SC R
IP
Andò, S., Garzanti, E., 2014. Raman spectroscopy in heavy-mineral studies. In: Scott, R.A., Smyth. H.R., Morton, A.C., Richardson, N. (Eds.), Sediment provenance studies in hydrocarbon exploration and production. Geological Society London Special Publication 386, 395-412. Andò, S., Vignola, P., Garzanti, E., 2011. Raman counting: a new method to determine provenance of silt. Rendiconti Lincei 22, 327-347.
NU
Andò, S., Garzanti, E., Padoan, M., Limonta, M., 2012. Corrosion of heavy minerals during weathering and diagenesis: a catalogue for optical analysis. Sedimentary Geology 280, 165-178.
MA
Bangs Rooney, C., Basu, A., 1994. Provenance analysis of muddy sandstones. Journal of Sedimentary Research 64, 2-7. Basu, A., 1985. Influence of climate and relief on compositions of sands released at source areas. In: Zuffa, G.G. (Ed.), Provenance of arenites. Reidel, Dordrecht, NATO ASI Series 148, 1-18.
TE
D
Bateman, R.M., Catt, J.A., 2007. Provenance and palaeoenvironmental interpretation of superficial deposits, with particular reference to post-depositional modification of heavy-mineral assemblages. In: Mange, M.A., Wright, D.T. (Eds.), Heavy Minerals in Use. Elsevier, Amsterdam, Developments in Sedimentology Series 58, pp. 151-188.
CE P
Bates, T.F., Griffiths, J.C., 1971. Memorial of Paul Dimitri Krynine. The American Mineralogist 56, 690-698.
AC
Bhatia, M.R., 1983. Plate tectonics and geochemical composition of sandstones. The Journal of Geology 91, 611-627. Blatt, H. 1967. Provenance determinations and recycling of sediments. Journal of Sedimentary Petrology 37, 1031-1044. Blatt, H. 1985. Provenance studies and mudrocks. Journal of Sedimentary Petrology 55, 69-75. Bloemsma, M.R., Zabel, M., Stuut, J.B.W., Tjallingii, R., Collins, J.A., Weltje, G.J., 2012. Modelling the joint variability of grain size and chemical composition in sediments. Sedimentary Geology 280, 135-148. Bluck, B.J., Dempster, T.J., Aftalion, M., Haughton, P.D.W., Rogers, G., 2006. Geochronology of a granitoid boulder from the Corsewall Formation (Southern Uplands): implications for the evolution of southern Scotland. Scottish Journal of Geology 42, 29-35. Borges, J.L., 1960. Del rigor de la ciencia. In: El Hacedor, p. 103. Emece Editores, Buenos Aires. Borges, J.B., Huh, Y., Moon, S., Noh, H., 2008. Provenance and weathering control on river bed sediments of the eastern Tibetan Plateau and the Russian Far East. Chemical Geology 254, 52-72.
ACCEPTED MANUSCRIPT
21
Bouchez, J., Lupker, M., Gaillardet, J., France-Lanord, C., Maurice, L., 2011. How important is it to integrate riverine suspended sediment chemical composition with depth? Clues from Amazon River depth-profiles. Geochimica et Cosmochimica Acta 75,6955-6970.
IP
T
Carrapa, B., Di Giulio, A., 2001. The sedimentary record of the exhumation of a granitic intrusion into a collisional setting: the lower Gonfolite Group, Southern Alps, Italy. Sedimentary Geology 139, 217-228.
SC R
Carver, R.E., 1971. Heavy-mineral separation. In: Carver, R.E. (Ed.), Procedures in sedimentary petrology. Wiley, New York, pp. 427-452.
NU
Clift P.D., Lee J.I., Hildebrand P., Shimizu N., Layne G.D., Blusztajn J., Blum J.D., Garzanti E., Khan A.A., 2002. Nd and Pb isotope variability in the Indus River system: implications for sediment provenance and crustal heterogeneity in the Western Himalaya. Earth Planetary Science Letters 200, 91-106. Crook, K.A.W., 1960. Classification of arenites. American Journal of Science 258, 419-428.
D
MA
Crook, K.A.W. l974, Lithogenesis and geotectonics: The significance of compositional variation in flysch arenites (graywackes). In: Dott, R.H., Shaver, R.H. (Eds.), Modern and ancient geosynclinal sedimentation. Society of Economic Paleontologists and Mineralogists Special Publication l9, pp. 304-3l0. Cummins, W.A., 1962. The greywacke problem. Geological Journal 3, 51-72.
TE
Dewey, J.F., 2005. Orogeny can be very short. Proceedings of the National Academy of Sciences of the United States of America 102, 15286-15293.
CE P
Dickinson, W.R., 1970. Interpreting detrital modes of graywacke and arkose. Journal of Sedimentary Petrology 40, 695-707.
AC
Dickinson, W.R., 1985. Interpreting provenance relations from detrital modes of sandstones. In: Zuffa, G.G. (Ed.), Provenance of arenites. Reidel, Dordrecht, NATO ASI Series 148, pp. 333-361. Dickinson, W.R., 1988. Provenance and sediment dispersal in relation to paleotectonics and paleogeography of sedimentary basins. In: Kleinspehn, K.L., Paola, C. (Eds.), New perspectives in basin analysis. Berlin, Springer, pp. 3-25. Dickinson, W.R., 2003. The place and power of myth in geoscience: an associate editor‟s perspective. American Journal of Science 303, 856-864. Dickinson, W.R., 2008. Impact of differential zircon fertility of granitoid basement rocks in North America on age populations of detrital zircons and implications for granite petrogenesis. Earth and Planetary Science Letters 275, 80-92. Dickinson, W.R., Rich, E.I., 1972. Petrologic intervals and petrofacies in the Great Valley sequence, Sacramento Valley, California. Geological Society of America Bulletin 83, 3007-3024. Dickinson, W.R., Suczek, C.A., 1979. Plate tectonics and sandstone composition. American Association of Petroleum Geologists Bulletin 63, 2164-2172. Dickinson, W.R., Beard, L.S., Brakenridge, G.R., Erjavec, J.L., Ferguson, R.C., Inman, K.F., Knepp, R.A., Lindberg, F.A., Ryberg, P.T., 1983. Provenance of North American Phanerozoic sandstones in relation to tectonic setting. Geological Society of America Bulletin 93, 222-235.
ACCEPTED MANUSCRIPT
22
Dott, R.H., 1964. Wacke, graywacke and matrix - what approach to immature sandstone classification? Journal of Sedimentary Petrology 34, 625-632. Dott, R.H., 2003. The importance of eolian abrasion in supermature quartz sandstones and the paradox of weathering on vegetation-free landscapes. The Journal of Geology 111, 387-405.
IP
T
Dunkl, I., Frisch, W., Kuhlemann, J., Brügel, A., 2009. Pebble population dating as an additional tool for provenance studies - examples from the Eastern Alps. Geological Society, London Special Publication 324, 125-140.
SC R
Folk, R.L., 1954. The distinction between grain size and mineral composition in sedimentary-rock nomenclature. The Journal of Geology 62, 344-359. Folk, R.L., 1980. Petrology of sedimentary rocks. Hemphill‟s, Austin,184 p.
NU
Gabriel, K.R., 1971. The biplot graphic display of matrices with application to principal component analysis. Biometrika 58, 453-467.
MA
Galehouse, J.S., 1971. Point counting. In: Carver, R.E. (Ed.), Procedures in sedimentary petrology. Wiley, New York, pp. 385-407. Garçon, M., Chauvel, C., France-Lanord, C., Limonta, M., Garzanti, E., 2014. Which minerals control the Nd–Hf–Sr–Pb isotopic compositions of river sediments? Chemical Geology 364, 42-55.
TE
D
Garzanti, E., Andò, S., 2007a. Heavy-mineral concentration in modern sands: implications for provenance interpretation. In: Mange, M.,Wright, D. (Eds.), Heavy minerals in use. Elsevier, Amsterdam, Developments in Sedimentology Series 58, pp. 517-545.
CE P
Garzanti, E., Andò S., 2007b. Plate tectonics and heavy-mineral suites of modern sands. In: Mange, M.A., Wright, D.T. (Eds.), Heavy Minerals in Use. Elsevier, Amsterdam, Developments in Sedimentology Series 58, pp. 741-763.
AC
Garzanti, E., Resentini, A., 2015. Provenance control on chemical indices of weathering (Taiwan river sands). Sedimentary Geology, this volume. Garzanti, E., Vezzoli, G., 2003. A classification of metamorphic grains in sands based on their composition and grade. Journal of Sedimentary Research 73, 830-837. Garzanti, E., Critelli, S., Ingersoll, R.V., 1996. Paleogeographic and paleotectonic evolution of the Himalayan Range as reflected by detrital modes of Tertiary sandstones and modern sands (Indus transect, India and Pakistan). Geological Society of America Bulletin 108, 631-642. Garzanti, E., Andò, S., Scutellà, M., 2000. Actualistic ophiolite provenance: the Cyprus case. The Journal of Geology 108, 199-218. Garzanti, E., Vezzoli, G., Andò, S., Castiglioni G., 2001. Petrology of rifted-margin sand (Red Sea and Gulf of Aden, Yemen). The Journal of Geology 109, 277-297. Garzanti, E., Canclini, S., Moretti Foggia, F., Petrella, N., 2002a. Unraveling magmatic and orogenic provenances in modern sands: the back-arc side of the Apennine thrust-belt (Italy). Journal of Sedimentary Research 72, 2-17. Garzanti, E., Vezzoli, G., Andò, S., 2002b. Modern sand from obducted ophiolite belts (Oman, U.A.E.). The Journal of Geology 110, 371-391.
ACCEPTED MANUSCRIPT
23
Garzanti, E., Andò, S., Vezzoli, G., Dell‟Era, D., 2003. From rifted margins to foreland basins: investigating provenance and sediment dispersal across desert Arabia (Oman, UAE). Journal of Sedimentary Research 73, 572-588.
IP
T
Garzanti, E., Vezzoli, G., Lombardo, B., Andò, S., Mauri, E., Monguzzi, S., Russo, M., 2004. Collision-orogen provenance (Western and Central Alps): detrital signatures and unroofing trends. The Journal of Geology 112, 145-164.
SC R
Garzanti, E., Andò, S., Vezzoli, G., 2006. The continental crust as a source of sand (Southern Alps cross-section, Northern Italy). The Journal of Geology 114, 533–554.
NU
Garzanti, E., Doglioni, C., Vezzoli, G., Andò S., 2007. Orogenic Belts and orogenic sediment provenances. The Journal of Geology 115, 315-334. Garzanti E., Andò S., Vezzoli G., 2008, Settling equivalence of detrital minerals and grain-size dependence of sediment composition. Earth and Planetary Science Letters 273, 138-151.
MA
Garzanti, E., Andò, S., Vezzoli, G., 2009. Grain-size dependence of sediment composition and environmental bias in provenance studies. Earth and Planetary Science Letters 277, 422–432.
TE
D
Garzanti, E., Andò, S., France-Lanord, C., Vezzoli, G., Galy, V., Najman, Y., 2010a. Mineralogical and chemical variability of fluvial sediments. 1. Bedload sand (Ganga–Brahmaputra, Bangladesh). Earth and Planetary Science Letters 299, 368-381.
CE P
Garzanti, E., Resentini, A., Vezzoli, G., Andò, S., Malusà, M.G., Padoan, M., Paparella, P., 2010b. Detrital fingerprints of fossil continental-subduction Zones (axial belt provenance, European Alps). The Journal of Geology 118, 341-362.
AC
Garzanti, E., Andò, S., France-Lanord, C., Galy, V., Censi, P., Vignola, P., 2011. Mineralogical and chemical variability of fluvial sediments. 2. Suspended-load silt (Ganga–Brahmaputra, Bangladesh). Earth and Planetary Science Letters 302, 107-120. Garzanti, E., Resentini, A., Vezzoli, G., Andò, S., Malusà, M., Padoan, M., 2012. Forward compositional modelling of Alpine orogenic sediments. Sedimentary Geology 280, 149-164. Garzanti, E., Padoan, M., Andò, S., Resentini, A., Vezzoli, G., Lustrino, M., 2013a. Weathering and relative durability of detrital minerals in equatorial climate: sand petrology and geochemistry in the East African Rift. The Journal of Geology 121, 547-580. Garzanti, E., Limonta, M., Resentini, A., Bandopadhyay, P.C., Najman, Y., Andò, S., Vezzoli, G., 2013b. Sediment recycling at convergent plate margins (Indo-Burman Ranges and AndamanNicobar Ridge). Earth-Science Reviews 123, 113-132. Garzanti, E., Vermeesch, P., Andò, S., Vezzoli, G., Valagussa, M., Allen, K., Khadi, K.A., AlJuboury, I.A., 2013c. Provenance and recycling of Arabian desert sand. Earth-Science Reviews 120, 1-19. Garzanti. E., Vermeesch, P., Padoan, M., Resentini, A., Vezzoli, G., Andò, S., 2014. Provenance of passive-margin sand (southern Africa). The Journal of Geology 122, 17-42.
ACCEPTED MANUSCRIPT
24
Garzanti, E., Resentini, A., Andò, S., Vezzoli, G., Vermeesch, P., 2015a. Physical controls on sand composition and relative durability of detrital minerals during long-distance littoral and eolian transport (coastal Namibia). Sedimentology 62, 971-996.
T
Garzanti, E., Andò, S., Padoan, M., Vezzoli, G., El Kammar, A., 2015b. The Nile sediment system : processes and products. Quaternary Science Reviews, in press.
IP
Gazzi, P., 1965. On the heavy mineral zones in the geosyncline series, recent studies in the Northern Apennines, Italy. Journal of Sedimentary Petrology 35, 109-115.
SC R
Gazzi, P., 1966. Le arenarie del flysch sopracretaceo dell‟ Appennino modenese: correlazioni con il flysch di Monghidoro. Mineralogica Petrographica Acta 12, 69-97.
NU
Goldstein, S.J., Jacobsen, S.B., 1988. The Nd and Sr isotopic systematics of river water suspended material: implications for crustal evolution. Earth and Planetary Science Letters 87, 249-265.
MA
Handy, M.R., 1990. The exhumation of cross sections of the continental crust: structure, kinematics and rheology. In: Salisbury, M.H., Fountain, D.M. (Eds.), Exposed cross-sections of the continental crust. Kluwer, Amsterdam, pp. 485-507. Hinderer, M., 2012. From gullies to mountain belts: A review of sediment budgets at various scales. Sedimentary Geology 280, 21-59.
TE
D
Huckenholz, H.G., 1963. Mineral composition and texture in graywackes from the Harz Mountains (Germany) and in arkoses from the Auvergne (France). Journal of Sedimentary Petrology 33, 914918.
CE P
Ingersoll, R.V. 1983. Petrofacies and provenance of Late Mesozoic forearc basin, northern and central California. American Association of Petroleum Geologists Bulletin 67, 1125-1142. Ingersoll, R.V., 1990. Actualistic sandstone petrofacies: discriminating modern and ancient source rocks. Geology 18, 733-736.
AC
Ingersoll, R.V., Bullard, T.F., Ford, R.L., Grimm, J.P., Pickle, J.D., Sares, S.W., 1984. The effect of grain size on detrital modes: a test of the Gazzi-Dickinson point-counting method, Journal of Sedimentary Petrology 54, 103-116. Johnsson, M.J., 1993. The system controlling the composition of clastic sediments. In: Johnsson, M.J., Basu, A. (Eds.), Processes controlling the composition of clastic sediments, Geological Society of America Special Paper 284, pp. 1-19. Kahneman, D., 2011. Thinking, fast and slow. Penguin, London, 499 p. Krynine, P.D., 1948, The megascopic study and field classification of sedimentary rocks. The Journal of Geology 56, 130-165. Krynine, P.D., 1956. Alice in Graywackeland. Journal of Sedimentary Petrology 26, 188-189. Lawrence et al., 2011. Hydrodynamic fractionation of zircon age populations. Geological Society of America Bulletin 123, 295-305. Le Pera, E., Critelli, S., 1997, Sourceland controls on the composition of beach and fluvial sand of the northern Tyrrhenian coast of Calabria, Italy: implications for actualistic petrofacies. Sedimentary Geology 110, 81-97.
ACCEPTED MANUSCRIPT
25
Limonta, M., Garzanti, E., Resentini, A., Andò, S., Boni, M., Bechstädt, T., 2015. Multicyclic sediment transfer along and across convergent plate boundaries (Barbados, Lesser Antilles). Basin Research, 1-18, DOI: 10.1111/bre.12095.
T
Mack, G.H., 1984. Exceptions to the relationship between plate tectonics and sandstone composition. Journal of Sedimentary Petrology 54, 212-220.
IP
Malusà, M.G., Resentini, A., Garzanti, E., 2015. Hydraulic sorting and mineral fertility bias in detrital geochronology. Gondwana Research, submitted May 2015.
SC R
Mange, M.A., Maurer, H.F.W., 1992. Heavy minerals in colour. Chapman and Hall, London, 147 pp. Mange, M.A., Wright, D.T., 2007. Heavy Minerals in Use. Elsevier, Amsterdam, Developments in Sedimentology Series 58, 1283 p.
NU
Marsaglia, K.M., Ingersoll, R.V., 1992. Compositional trends in arc-related, deep-marine sand and sandstone: a reassessment of magmatic-arc provenance. Geological Society of America Bulletin 104, 1637-1649.
MA
McBride, E.F., Picard, D.M., 1987. Downstream changes in sand composition, roundness and gravel size in a short-headed, high-gradient stream, Northwestern Italy. Journal of Sedimentary Petrology 57, 1018-1026.
TE
D
McBride, E.F., 1963. A classification of common sandstones. Journal of Sedimentary Petrology 33, 664-669.
CE P
McBride, E.F., 1985. Diagenetic processes that affect provenance determinations in sandstones. In: Zuffa, G.G. (Ed.), Provenance of arenites. Reidel, Dordrecht, NATO ASI Series 148, pp. 95-113. Moecher, D.P., Samson, S.D., 2006. Differential zircon fertility of source terranes and natural bias in the detrital zircon record: implications for sedimentary provenance analysis. Earth and Planetary Science Letters 247, 252-266.
AC
Molinaroli, E., Blom, M., Basu, A., 1991. Methods of provenance determination tested with discriminant function analysis. Journal of Sedimentary Research 61, 900-908. Morton, A.C., 1985. Heavy minerals in provenance studies. In: Zuffa, G.G. (Ed.), Provenance of arenites. Reidel, Dordrecht, NATO ASI Series 148, pp. 249-277. Morton, A.C., 2012. Value of heavy minerals in sediments and sedimentary rocks for provenance, transport history and stratigraphic correlation. In: Sylvester, P. (Ed.), Quantitative Mineralogy and Microanalysis of Sediments and Sedimentary Rocks. Mineralogical Association of Canada Short Course Series 42, pp. 133-165. Morton, A.C., Hallsworth, C., 2007. Stability of detrital heavy minerals during burial diagenesis. In: Mange, M.A., Wright, D.T. (Eds.), Heavy Minerals in Use. Elsevier, Amsterdam, Developments in Sedimentology Series 58, pp. 215-245. Murchison, R.I., 1854. Siluria. Murray, London, 523 p. Najman, Y., 2006. The detrital record of orogenesis: a review of approaches and techniques used in the Himalayan sedimentary basins. Earth-Science Reviews 74, 1-72. Nesbitt, H.W., Young, G.M., 1982. Early Proterozoic climates and plate motions inferred from major element chemistry of lutites. Nature 299, 715-717.
ACCEPTED MANUSCRIPT
26
Nesbitt, H.W., Young, G.M., 1996. Petrogenesis of sediments in the absence of chemical weathering: effects of abrasion and sorting on bulk composition and mineralogy. Sedimentology 43, 341-358. Padoan, M., Garzanti, E., Haravan, Y., Villa, I.M., 2011. Tracing Nile sediment sources by Sr and Nd isotope signatures (Uganda, Ethiopia, Sudan). Geochimica et Cosmochimica Acta 75, 3627-3644.
IP
T
Paola, C., Leeder, M., 2011. Environmental dynamics: simplicity versus complexity. Nature 469, 3839.
SC R
Pettijohn, F.J., 1948. A preface to the classification of the sedimentary rocks. The Journal of Geology 56, 112-117. Pettijohn, F.J., 1954. Classification of sandstones. The Journal of Geology 62, 360-365.
NU
Pearce, J., Cann, J., 1973. Tectonic setting of basic volcanic rocks determined using trace element analyses. Earth and Planetary Science Letters 19, 290-300.
MA
Potter, P.E., Huh, Y., Edmond, J.M., 2001. Deep-freeze petrology of Lena River sand, Siberia. Geology 29, 999-1002. Price, J.R., Velbel, M.A., 2003. Chemical weathering indices applied to weathering profiles developed on heterogeneous felsic metamorphic parent rocks. Chemical Geology 202, 397-416.
D
Roser, B.P., Korsch, R.J., 1986. Determination of tectonic setting of sandstone-mudstone suites using SiO2 content and K2O/Na2O ratio. The Journal of Geology 94, 635-650.
TE
Rubey, W.W., 1933. The size-distribution of heavy minerals within a water-laid sandstone. Journal of Sedimentary Petrology 3, 3-29.
CE P
Shao, J., Yang, S., Li, C., 2012. Chemical indices (CIA and WIP) as proxies for integrated chemical weathering in China: inferences from analysis of fluvial sediments. Sedimentary Geology 265, 110120.
AC
Spalla, M.I., Zanoni, D., Gosso, G., Zucali, M., 2009. Deciphering the geologic memory of a Permian conglomerate of the Southern Alps by pebble P–T estimates. International Journal of Earth Sciences 98, 203-226. Suttner, L.J., 1974. Sedimentary petrographic provinces: an evaluation. In: Ross, C.A. (ed.), Paleogeographic provinces and provinciality. Society of Economic Paleontologists and Mineralogists Special Publication 21, pp. 75-84. Suttner, L.J., Basu, A., 1985. The effect of grain size on detrital modes: a test of the Gazzi-Dickinson point-counting method: discussion. Journal of Sedimentary Petrology 55, 616-618. Suttner, L.J., Basu, A., Mack, G.H., 1981. Climate and the origin of quartz arenites. Journal of Sedimentary Petrology 51, 1235-1246. Taylor, S.R., McLennan, S.M., 1995. The geochemical evolution of the continental crust. Reviews of Geophysics 33, 241-265. Velbel, M.A., 2007. Surface textures and dissolution processes of heavy minerals in the sedimentary cycle: examples from pyroxenes and amphiboles. In: Mange, M.A., Wright, D.T. (Eds.), Heavy Minerals in Use. Elsevier, Amsterdam, Developments in Sedimentology Series 58, pp. 112-150.
ACCEPTED MANUSCRIPT
27
Vermeesch, P., Avigad, D., McWilliams, M.O., 2009. 500 m.y. of thermal history elucidated by multimethod detrital thermochronology of North Gondwana Cambrian sandstone (Eilat area, Israel). Geological Society of America Bulletin 121, 1204-1216.
T
von Eynatten, H., Dunkl, I., 2012. Assessing the sediment factory: the role of single grain analysis. Earth-Science Reviews 115, 97-120.
SC R
IP
von Eynatten, H., Barcelò-Vidal, C., Pawlowsky-Glahn, V., 2003. Composition and discrimination of sandstones: a statistical evaluation of different analytical methods. Journal of Sedimentary Research 73, 47–57. von Eynatten, H., Tolosana-Delgado, R., Karius, V., 2012. Sediment generation in modern glacial settings: grain-size and source-rock control on sediment composition. Sedimentary Geology 280, 80-92.
MA
NU
von Eynatten, H., Tolosana-Delgado, R., Karius, V., Bachmann, K., Caracciolo, L., 2015. Sediment generation in humid Mediterranean setting: grain-size and source-rock control on sediment geochemistry and mineralogy (Sila Massif, Calabria). Sedimentary Geology, this volume. Weltje, G.J., 2004. A quantitative approach to capturing the compositional variability of modern sands. Sedimentary Geology 171, 59-77.
TE
D
Weltje, G.J., 2006. Ternary sandstone composition and provenance: an evaluation of the „Dickinson model‟. In: Buccianti, A., Mateu-Figueras, G., Pawlowsky-Glahn, V. (Eds.), Compositional data analysis: from theory to practice. Geological Society of London Special Publications 264, pp. 611627.
CE P
Weltje, G.J., 2012. Quantitative models of sediment generation and provenance: State of the art and future developments. Sedimentary Geology 280, 4-20. Weltje, G.J., von Eynatten, H., 2004. Quantitative provenance analysis of sediments: an introduction. Sedimentary Geology 171, 1-11.
AC
Weltje, G.J., Bloemsma, M.R., Boekhout, S.G., Garzanti, E., 2015. Compositional variation of clastic sediments resulting from size-shape-density sorting of sediments in suspension: a process-based statistical model for quantitative provenance analysis. Sedimentary Geology, this volume. Winchester, J.A., Floyd, P.A., 1977. Geochemical discrimination of different magma series and their differentiation products using immobile elements. Chemical Geology 20, 325-343. Wittgenstein, L., 1974. On Certainty. Basil Blackwell, Oxford, #1-676. Zuffa, G.G., 1985. Optical analyses of arenites: influence of methodology on compositional results. In: Zuffa, G.G. (Ed.), Provenance of arenites. Reidel, Dordrecht, NATO ASI Series 148, pp. 165-189. Zuffa, G.G., 1987. Unravelling hinterland and offshore palaeogeography from deep-water arenites. In: Leggett, J.K., Zuffa, G.G. (Eds.), Marine clastic sedimentology. Concepts and case studies. Graham and Trotman, London, pp. 39-61.
ACCEPTED MANUSCRIPT
28
FIGURE CAPTIONS
Figure 1. Simple descriptive petrographic classification for sands and sandstones. Following the
T
nomenclature scheme introduced by Crook (1960) and Dickinson (1970), sands and sandstones are
IP
classified according to their main components exceeding 10% QFL. An adjective reflecting the dominant rock-fragment type may be added freely (e.g., volcaniclastic, sedimentaclastic,
SC R
metamorphiclastic, ultramaficlastic; Ingersoll 1983). The lithic pole L should include carbonate and chert grains. Shown to the upper right is the same diagram in logratio space (Weltje, 2006).
NU
Figure 2. Composition of modern sands generated along divergent plate margins (after Garzanti et al., 2001, 2013a, 2013c, 2014, 2015 plus unpublished data; the lithic pole L includes carbonate and
MA
chert grains). These four datasets, covering passive margins sands around southern Africa, much of the East Africa and Red Sea-Levant rift system from Botswana to Syria and the entire Nile catchment, demonstrate that detrital modes cannot be used as a proxy for plate-tectonic setting. A)
D
Anorogenic and orogenic sands overlap extensively, and cannot be discriminated by their QFL
TE
modes. B) Sands of Dissected Continental Block provenance are controlled strictly by climaterelated weathering. Sands of Undissected to Transitional Continental Block provenance may plot in
CE P
the Recycled Orogen field. The Anorogenic Volcanic and Magmatic Arc fields overlap completely. C) Quartzo-feldspatho-lithic suspended load derived from Ethiopian volcanic highlands mixes with feldspatho-quartzose bedload shed by crystalline basement exposed in the Blue Nile catchment; quartzose White Nile sands are coarser than volcaniclastic Atbara sediments. Main Nile sediments
AC
thus overlap all three Dickinson's (1985) provenance subdivisions, depending on their grain size: silty levees plot in the Magmatic Arc field (MA), fine bar sands in the Recycled Orogen field (RO), and medium bar sands in the Continental Block field (CB). D) Sediments eroded from rift shoulders are mixtures in virtually all proportions of detritus from crystalline basement (Dissected Continental Block provenance), sedimentary covers (Undissected Continental Block provenance) and volcanic fields (Anorogenic Volcanic provenance).
Figure 3. Composition of modern sands generated along convergent plate margins (after Garzanti et al., 2007; the lithic pole L includes carbonate and chert grains). The five primary types of orogenic provenance overlap widely on the QFL plot, which forbids its uncritical use as an oracle. The key information lies in the rock fragments. A quick glance at the microscope is sufficient to discriminate volcaniclastic to plutoniclastic sands of Magmatic Arc provenance, basalticlastic to ultramaficlastic sands of Ophiolite provenance, metamorphiclastic sands of Axial Belt provenance, sedimentaclastic
ACCEPTED MANUSCRIPT
29
to basementaclastic sands of Continental Block provenance, and lithic to quartzose sands of Recycled Clastic provenance. Grey arrows highlight unroofing trends for each provenance. Fields after Dickinson (1985): MA= Magmatic Arc; CB= Continental Block; RO= Recycled Orogen.
AC
CE P
TE
D
MA
NU
SC R
IP
T
Coloured scale bars are all 250 m.
AC
CE P
TE
D
MA
NU
SC R
IP
T
ACCEPTED MANUSCRIPT
30
AC
CE P
TE
D
MA
NU
SC R
IP
T
ACCEPTED MANUSCRIPT
31
AC
CE P
TE
D
MA
NU
SC R
IP
T
ACCEPTED MANUSCRIPT
32