You are on page 1of 55

Accepted Manuscript

Petrographic classification of sand and sandstone

Eduardo Garzanti

PII: S0012-8252(18)30606-8
DOI: https://doi.org/10.1016/j.earscirev.2018.12.014
Reference: EARTH 2754
To appear in: Earth-Science Reviews
Received date: 18 October 2018
Revised date: 16 December 2018
Accepted date: 18 December 2018

Please cite this article as: Eduardo Garzanti , Petrographic classification of sand and
sandstone. Earth (2018), https://doi.org/10.1016/j.earscirev.2018.12.014

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.
ACCEPTED MANUSCRIPT
1

PETROGRAPHIC CLASSIFICATION OF SAND AND SANDSTONE

Eduardo Garzanti eduardo.garzanti@unimib.it

Laboratory of Provenance Studies, Department of Earth and Environmental Sciences, University of

PT
Milano-Bicocca, Milano (Italy)
Tel: +39 02 64482088, Email:

RI
SC
NU
ABSTRACT

Descriptive petrographic classifications of sand and sandstone proposed more than half a century
MA

ago are still in use, although they were formulated at a time when depositional and post-depositional

sedimentary processes were poorly understood, and before the relationships between tectonics and
ED

sedimentation could be interpreted in modern plate-tectonic terms. As a consequence, too many

scientific articles and technical reports are still encumbered with obsolete concepts, graphical tools,
T

and ambiguous terminology that make sediment descriptions awkward and misleading. A
EP

renovation that treasures the legacy of the pioneers is required.


C

The descriptive petrographic classification of sand and sandstone proposed in this paper is based on
AC

the quasi-universally used Gazzi-Dickinson point-counting method, and simply translates into

words ternary compositions of quartz, feldspar, and lithic fragments without introducing any new

names. The classic QFL plot is subdivided into 15 fields - labelled by adjectives introduced long

ago by K.A.W. Crook and endorsed by W.R. Dickinson and more recently by G.J. Weltje - which

reflect relative abundances of the three main framework components (provided they exceed

10%QFL). According to standard use, the less abundant component goes first, the more abundant

last (e.g., litho-feldspatho-quartzose composition translates into Q > F > L >10%QFL). For lithic-

rich sand and sandstone, information on the prevailing rock fragment type can be added by an
ACCEPTED MANUSCRIPT
2

additional free adjective (e.g., metamorphiclastic, carbonaticlastic), as proposed long ago by R.V.

Ingersoll. For lithic-poor feldspatho-quartzose and quartzose sand and sandstone, further formal

subdivisions are proposed, thus reaching a total of 18 compositional fields overall. Modern sand

known to be derived from different source rocks and found in major world’s rivers, deserts, and

deep-sea fans fits in the pigeonholes defined by the relative abundance of quartz, feldspar, and lithic

fragments.

PT
The aim of this classification is to restore directness in sandstone petrology, and to avoid

ambiguities generated in the past by making reference to badly defined archetypes, such as

RI
greywacke or arkose, thus confusing petrographic composition with subjective considerations about

SC
plate-tectonic setting, texture, hydraulic behaviour, mechanical durability, or chemical durability in

the illusion that a classification could be genetic at the same time as descriptive.
NU
“You ask what is the use of classification, arrangement, systematization?
I answer you: order and simplification are the first steps toward the
mastery of a subject — the actual enemy is the unknown.”
MA

(Thomas Mann, The Magic Mountain, Encyclopaedic)


ED

Keywords
Sandstone petrology; Gazzi-Dickinson method; QFL diagram; Sediment provenance; Plate-tectonic
T

setting; Texture and composition; Greywacke and arkose; Rock fragments; Chert, carbonate, and
EP

evaporite grains; Classification of modern river, turbidite, and desert sand.

1. INTRODUCTION
C
AC

The petrographic study of sedimentary archives is one of the many keys to deciphering geological

history. The road is, however, long and winding, and a firm methodological approach is essential to

staying on track. This article reviews operational and conceptual flaws of traditional classification

criteria and suggests how to describe and classify in a clear and exhaustive way the composition of

sand or sandstone samples in a scientific article or technical report. The straightforward scheme

proposed here  which focuses strictly on petrographic composition, expressly neglecting texture,
ACCEPTED MANUSCRIPT
3

hydraulic behaviour, resistance to weathering, transport, or diagenesis  is intended to be efficient

and efficacious at the same time.

After the invention of thin-section petrography by Sorby (1880), sedimentary-petrology studies

started to flourish in the first half of the twentieth century, and culminated with sandstone

classifications proposed in numbers between the late 1940s and the early 1960s (as summarized in

Klein, 1960; Okada, 1971; Scholle, 1979). These classification schemes (Fig. 1; Fig. 2) inevitably

PT
suffered from the level of information available at the time about sedimentary and geodynamic

RI
processes. Investigations on diagenetic transformations were limited, and the nature of interstitial

SC
phyllosilicates in sandstone consequently poorly understood. Before advent of the plate-tectonic

theory, relationships between tectonics and sedimentation were still immersed in a panorama full of
NU
mythological entities, including the large geosyncline clan (Krynine, 1948; Kay, 1951; Folk, 1968).

Irruption of the new geodynamic paradigm in the 1960s and 1970s revolutionized the petrological
MA

study of terrigenous rocks as well, leading to novel ideas embodied in the genetic sandstone-

classification scheme of Dickinson (1985; Dickinson and Suczek, 1979) that has dominated the
ED

scene since then. The enthusiasm generated by such a major conceptual step forward, however,

created the illusion that sediments generated in different plate-tectonic settings should inevitably
T
EP

plot in separate fields within a QFL diagram. The uncritical use of such a simple graphic tool as a

passkey to paleogeodynamic interpretation has resisted the objections manifested in subsequent


C

years (e.g., Mack, 1984; Molinaroli et al., 1991; Weltje, 2006), ending up as a ready-made approach
AC

that has all too often contributed to sterilization of petrological research.

This conservative attitude has persisted. Even in recent scientific articles, it is common to find

cumbersome petrographic descriptions based on obsolete classification schemes or awkward terms

such as arkose or greywacke, the use of which has been contested since their early introduction two

centuries ago. In too many cases, genetic interpretations are still based on the belief that numerical

parameters readily obtained by petrographic analysis may open an easy way to the reconstruction of

paleogeographic and paleogeodynamic scenarios. A fundamental misunderstanding rooted in the


ACCEPTED MANUSCRIPT
4

past (Pettijohn, 1948 p.113; Rodgers, 1950 p.299) is that a classification could, and even should be

at the same time both descriptive and genetic (i.e., objective and subjective). Sediment mineralogy,

however, reflects the multiple superposed effects of numerous controlling factors, including source-

rock lithology, climate, and tectonic activity, together with diverse physical and chemical processes

affecting detritus through one or more sedimentary cycles. Complex equations with many

unknowns cannot be solved in one leap.

PT
2. WHY A CLASSIFICATION SHOULD NOT BE GENETIC

RI
SC
Whenever we find ourselves baffled by the variety and complexity of nature’s products and

phenomena, we may start to make comparisons, look for similarities and differences, and finally
NU
combat chaos by dividing objects into categories and giving them a name. A classification is a

simple artificial means to impose order upon the real world. This process leads to the formulation of
MA

a language that, when widely agreed upon, allows us to exchange information. Words, however, do

not only technically describe objects, but also create suggestions. Not differently from myth and
ED

religion, scientific theories grow with new words, the spell of which may help to conceal the
T

unbridgeable distance separating our models from truth. In the geosciences, evocative words have
EP

always been used to depict processes and scenarios which, once meeting wide consensus, proved

easy to believe and difficult to depart from (Dickinson, 2003). In order to make diligent use of
C
AC

reason, digest observations, and progress on the infinite stairway to knowledge, however, we need

to focus on naked phenomena, stripped of rhetoric, superposed scent of incense, and special effects

of any kind. Sharp reasoning needs clear words and concepts devoid of ambiguities and implicit

assumptions.

A classification, therefore, should be no more than a basic tool, an attempt to communicate reality

in a direct way. An efficient descriptive language cannot substitute for understanding, but may

represent the first step on the path toward understanding. Hoping that the goal can be reached

without making the journey is laziness. Believing that enlightenment may come in a moment is
ACCEPTED MANUSCRIPT
5

illusory. The commandment “genesis must and does permeate our classification” (Pettijohn, 1948

p.113) or the promise that plotting a point into a QFL diagram is sufficient to reveal a geodynamic

context (Dickinson and Suczek, 1979) are traps into which we should not fall.

3. WHY A QFL PLOT CANNOT REVEAL GEODYNAMIC SETTING

PT
Plate tectonics is the best paradigm at our disposal to describe geological processes at planetary

scale. This does not mean that full understanding has been reached, and several key questions

RI
remain unanswered about the processes that shape the face of our planet. We are still debating when

SC
and how plate tectonics began on Earth, why plates are moving, what drives oceanic and continental

subduction, whether plates breakoff at depth, what forces cause subsidence in orogen-related basins,
NU
or what triggers gigantic outbursts of magmas (e.g., Hamilton, 2011; Korenaga, 2013; Doglioni and
MA

Panza, 2015; Garzanti et al., 2017a). Tectonic processes are reflected in sediment composition, and

sedimentary petrology thus represents one fruitful way to extrapolate knowledge acquired in
ED

modern settings to reconstruct tectonic evolution throughout the geological past. Using sediment

composition as a key to paleogeodynamics is the fundamental idea underlying the work of W.R.
T

Dickinson and co-workers (e.g., Dickinson and Suczek, 1979; Ingersoll and Suczek, 1979), an idea
EP

that stands as valid as ever. Significant parts of the operative procedure, however, need to be
C

revised.
AC

As shown by Molinaroli et al. (1991) and Weltje (2006), the uncritical use of Dickinson’s plots is

bound to meet with limited success, and not only because relevant factors such as grain-size control,

sampling scale, and environmental or diagenetic bias are neglected (Ingersoll, 1990; Johnsson,

1993; Weltje and von Eynatten, 2004; Basu, 2017) but also and more fundamentally because some

implicit assumptions about sediment generation embodied in this model are invalid. These include

the tenet that vast areas covered by continental flood basalts cannot supply large amounts of detritus

to sedimentary basins, even though anorogenic volcanic signatures characterize vast


ACCEPTED MANUSCRIPT
6

transcontinental river systems such as the Nile, the Orange, or to a lesser extent the Yang Tze

(Garzanti et al., 2006a, 2012; Vezzoli et al., 2016). Flawed is also the hope that petrographic

analysis alone could discriminate between detritus shed from neometamorphic domains forming the

axial core of young mountain belts versus paleometamorphic detritus derived from old orogenic

roots exposed in cratonic shields or uplifted on the shoulders of continental rifts. In the lack of

information from detrital geochronology and geochemistry, orogenic sandstone shed from thrust

PT
belts and magmatic arcs cannot be safely distinguished from anorogenic sandstone derived from

RI
continental blocks and flood basalts (e,g., Garzanti et al., 2015b). Framework petrography cannot

SC
tell us whether source rocks are allochthonous or autochthonous. Therefore, orogenic and

anorogenic provenances cannot be discriminated without additional geological information, and


NU
geodynamic setting cannot be univocally inferred from detrital modes of sandstone alone (Garzanti,

2016).
MA

4. WHY A PETROGRAPHIC CLASSIFICATION SHOULD NOT CONSIDER TEXTURE


ED

Texture and composition are independent variables. Texture of clastic sediments is primarily
T

defined by parameters measuring the characteristics of the grain-size distribution in progressive


EP

detail, i.e., average (mean, median, mode), uniformity (sorting), asymmetry (skewness), and
C

peakedness (kurtosis; Folk, 1966), whereas composition is defined by the relative percentages of
AC

framework components, i.e., single minerals and rock fragments. The repeated attempts to combine

in a single classification scheme compositional, textural, and even hydraulic properties at the same

time (provenance, maturity, and fluidity factors of Pettijohn, 1954) has created much confusion

since the dawn of sandstone petrography (Table 1; Klein, 1963 p.569-570) and still reverberates in

the appearance of ambiguous obsolete names in the recent geological literature.

The emblematic example is the archaic term greywacke, used originally by miners to describe hand

specimens and successively adopted in the geological literature to indicate a generic lithic-rich
ACCEPTED MANUSCRIPT
7

composition as well as colour (grey), texture (poorly sorted, supposedly matrix-rich), and

depositional environment (turbidite). The crux of ambiguity stems from putting composition and

texture in the same basket (Fig. 2), often implying that “immature” composition (i.e., low quartz)

should necessarily combine with “immature” texture (i.e., poor sorting). For instance, in Pettijohn

(1954 p.362) we read: “compositional maturity is rarely attained without a corresponding

achievement of textural maturity”. Many sandstone petrographers thus shared G.H. Packham’s view

PT
(1954) that a “greywacke suite” deposited by turbidity currents could be distinguished from an

RI
“arkose-quartzose sandstone suite” deposited by traction currents (McBride 1963; Crook 1974), a

dichotomy that has no basis (Okada, 1966; Garzanti, 2017). The tangle tightened since “clay

SC
matrix” was chosen as an end-member of sandstone composition in most formal classification
NU
schemes proposed in the 1950s (e.g., Pettijohn 1949; Dapples et al. 1953; Gilbert 1954; Packham

1954; Bokman 1955; Crook 1960). Later on, several authors emphasized the rarity of modern sand
MA

characterized by very poor sorting (Cummins, 1962; Whetten, 1969), and concluded that the

supposedly peculiar aspect of greywackes was not original, but produced during diagenesis by a
ED

variety of processes including plastic deformation of altered volcanic rock fragments or other soft

clasts of either extrabasinal or intrabasinal origin to create what has been termed pseudomatrix
T
EP

(Dickinson, 1970; Whetten and Hawkins, 1970; Galloway, 1974). The “greywacke problem” was

solved.
C
AC

4.1. Why greywackes should be forgotten

The term greywacke (grauwacke in German, graywacke in American English) was formally

introduced in 1785 by mine director F.W.H. von Trebra, adopted by A.G. Werner (1787 p.18) and

defined as “quartz breccias with flakes of micas and fragments of chert or sandstones in a cement of

clay” (Lasius, 1789). The type greywackes are upper Paleozoic deep-water sandstones of the Hartz

mountains that contain quartz as well as abundant feldspar and diverse plutonic, volcanic,

metamorphic, and sedimentary rock fragments (Helmbold and van Houten, 1958; Huckenholz,
ACCEPTED MANUSCRIPT
8

1963). The term has generated confusion and has been harshly criticized since its early introduction

(Mawe, 1818 p.92; Sedgwick and Murchison, 1839 p.260; Murchison, 1854 p.359; Krynine, 1956;

Boswell, 1960; Okada, 1971), until Folk (1968 p.128) conclusively pointed out that greywacke is

nothing else than “a very hard, ugly, dirty, dark rock that you can't tell much about in the field”.

4.2. Why arkoses should be forgotten

PT
The term arkose was introduced by Brongniart (1826) and originally defined as sandstone

RI
containing either more quartz than feldspar (arkose commune) or more feldspar than quartz (arkose

SC
granitoide). The type arkose derived from the Massif Central in France, however, may contain a

variety of igneous (granite), sedimentary (sandstone), and metamorphic (quartzite) rock fragments,
NU
largely overlapping the type greywacke in both texture and composition (Huckenholz, 1963; Dott,

1964 p.626; Dickinson, 1970 p.697). As for greywacke, sub-greywacke, high-rank greywacke and
MA

low-rank greywacke, the compositions of arkose (or “arkosite”), impure arkose, arkosic arenite,

arkosic wacke, or subarkose have been defined through time with inconsistent boundaries (Fig. 1;
ED

Fig. 2; Oriel, 1949; Scholle, 1979). Consequently, arkose and subarkose are at best imprecise
T

synonyms of generally but not necessarily lithic-poor, feldspar-rich and feldspar-bearing


EP

sandstones. As R.H. Dott (1964 p.625) put it: “the name arkose itself has little descriptive merit;

feldspathic sandstone would seem far more useful”.


C
AC

5. PETROGRAPHIC CLASSIFICATION OF SAND AND SANDSTONE

Terrigenous sediments are random assemblages of monomineralic and polymineralic particles

derived from any type of source rock, garbage cans that potentially host any kind of extrabasinal

and intrabasinal, larger and smaller, platy and equant, low-density and high-density, carbonate and

non-carbonate, felsic and mafic silicate grains. Making order is necessary.


ACCEPTED MANUSCRIPT
9

By the simple procedure presented here, a composite descriptive label is attached to sand or

sandstone samples based on their petrographic composition (Fig. 3). Because communicative

efficiency makes a scientific article or technical report shorter and clearer, the aim is to transmit the

richest information with the fewest words. If not widely agreed upon and adopted, however, a new

terminology only introduces unwanted noise. Taking advantage of already existing methods and

familiar terms, provided they are appropriate, is sensible, economic, and may facilitate acceptance

PT
by the community. For this reason, the proposed classification: a) is based on a consolidated

RI
analytical technique (i.e., Gazzi-Dickinson method; Ingersoll et al., 1984; Zuffa, 1985); b) makes

SC
use of fundamental parameters and of a universally recognized graphical display (i.e., QFL plot); c)

is expressed by readily understood compositional terms that follow a straightforward scheme


NU
proposed and endorsed by authoritative researchers since the 1960s.
MA

5.1. The Gazzi-Dickinson method

Gazzi (1966 p.73-74) started from the obvious consideration that coarse-grained rock fragments
ED

inevitably tend to be more abundant in coarser-grained samples. Therefore, in order to obtain


T

comparable quantitative data from samples of different grain size, he proposed that minerals
EP

occurring within rock fragments and exceeding 30 m in size – the conventional boundary between

frictional grains and cohesive matrix as well as the thickness of a standard thin section (Spencer,
C

1963; Dott, 1964 p.630-631) – should be reunited in the dataset with single detrital minerals of the
AC

same type.

A similar, albeit simplified version of this operational procedure was proposed independently a few

years later by Dickinson (1970), who distinguished formally between aphanite lithic fragments (L)

to be counted as such, and microphanerite rock fragments (R) to be assigned according to the

mineral beneath the cross-hair. The boundary between fine-grained aphanites and coarse-grained

phanerites was chosen instead as 62.5 m (i.e., the lower limit of the sand range), as agreed by P.

Gazzi’s former student G.G. Zuffa (1980 p.27, 1985). As a partial solution to this problem in case
ACCEPTED MANUSCRIPT
10

of data collection with the traditional QFR method, in which rock fragments are counted as such,

(Suttner et al., 1981), Folk (1980 p.127) proposed to group coarse-grained granitoid and granitoid-

gneiss grains with feldspar in pole F, and to assign only fine-grained volcanic, metamorphic, and

sedimentary rock fragments to pole R (Fig. 1).

The Gazzi-Dickinson method was eventually formalized by Ingersoll et al. (1984) and successively

widely accepted by most sandstone petrographers under the belief that “use of the method minimizes

PT
variation of composition with grain size, thus eliminating the need for sieving and multiple counts

RI
of different size fractions”, which is overly optimistic (Garzanti et al., 2006a p.331). The rectified

statement found a dozen lines below in Ingersoll et al. (1984 p.103) reads: “there are two reasons

SC
for variation of modal composition with grain size: 1) the breakage of fragments into constituent
NU
grains, and 2) actual mineralogic variation with grain size. The Gazzi-Dickinson method

successfully eliminates the first source of compositional grain-size dependency. No point-counting


MA

method eliminates the second source.”


ED

5.2. The QFL triangle


T

A consensus has long been reached among sedimentary petrographers that, as a first approximation,
EP

sand and sandstone can be considered as ternary mixtures of three principal components: quartz (Q;

most common mineral species in the Earth’s crust), feldspars (F; most common mineral group in the
C

Earth’s crust), and monomineralic or polymineralic rock or lithic fragments (R or L, respectively).


AC

Other components are generally less abundant, considered as accessories, and neglected for

classification purposes not only to reduce complexity but also to come closer to a transport-

invariant measure of sediment composition (Weltje, 2004). Most accessories have peculiar

hydraulic behaviour because of their distinct shapes (e.g., slow-settling platy phyllosilicates) or

densities (e.g., fast-settling dense minerals), and can be strongly concentrated or depleted locally in

sedimentary deposits, thus primarily reflecting physical processes in depositional environments

(Garzanti et al., 2008, 2009). Intrabasinal grains, which in sediment samples may be mixed in any
ACCEPTED MANUSCRIPT
11

proportion with terrigenous extrabasinal detritus derived from erosion of source rocks, must be

considered independently for classification purposes (fig. 3 in Zuffa, 1980; fig. 1 in Garzanti, 1991).

Once major framework components are reduced to three, sand and sandstone can be readily

classified making use of the classical QFL triangular diagram, in which each point represents a

ternary composition obtained by the Gazzi-Dickinson point-counting method. A most precise way

to convey compositional information would be represented by the relative numerical proportions

PT
among the three major components (e.g., with a format like Q 45 F18 L37 ; Dickinson, 1970). Although

RI
appropriate for tabulated data, however, an analytical script such as this singles out each object of

SC
the suite, and a classification of singles is no classification at all.
NU
5.3. Nomenclature

Most traditional classifications of sand and sandstone are based on the quartz-feldspar-lithics (QFL)
MA

or quartz-feldspar-rock fragments (QFR) triangle, subdivided into several fields (generally 5 to 10),

each labelled differently and delimited by necessarily arbitrary conventional boundaries (Fig.1; Fig.
ED

2). The first step of the procedure proposed here is the same as in Weltje (2006 p.82): by tracing

“three lines from each of the vertices towards the middle of the opposite sides, i.e. lines along which
T
EP

the abundance of one component equals that of another” the QFL space is subdivided in 6 equal

right-angled triangles labelled as litho-feldspathic, feldspatho-lithic, quartzo-lithic, litho-quartzose,


C

feldspatho-quartzose, and quartzo-feldspathic (Fig. 3B). The adjectives feldspathic (F/L > 3), litho-
AC

feldspathic (3 > F/L > 1), feldspatho-lithic (3 > L/F > 1) and lithic (L/F > 3) were originally

proposed by Crook (1960 p.425), and considered by Dickinson (1970 p.697) to “permit adequate

discrimination without confusion”.

G.J. Weltje’s straightforward classification, however, considers only the ratio between two of the

three major components, and conveys no information on the abundance of the third. Moreover, very

different compositions (e.g., Q 99 F1 L0 and Q 35 F33 L32 ) would plot in the same field. Additional fields

are needed to increase classification efficiency and discrimination power. A simple effective way is
ACCEPTED MANUSCRIPT
12

to trace another three lines corresponding to a 10% relative content of each major component, thus

obtaining another 9 fields (15 overall; Fig. 3C). The six central fields, where all three major

components exceed 10%, are equal right-angled triangles labelled as quartzo-litho-feldspathic

(qLF), quartzo-feldspatho-lithic (qFL), feldspatho-quartzo-lithic (fQL), feldspatho-litho-quartzose

(fLQ), litho-feldspatho-quartzose (lFQ), and litho-quartzo-feldspathic (lQF). The six trapezoidal

fields along the three legs of the QFL triangle, where one major component does not exceed 10%,

PT
are labelled as litho-feldspathic (LF), feldspatho-lithic (FL), quartzo-lithic (QL), litho-quartzose

RI
(LQ), feldspatho-quartzose (FQ), and quartzo-feldspathic (QF). The three rhomboidal fields at the

SC
apices, where two out of the three major components do not exceed 10%, are thus simply labelled

lithic (L), feldspathic (F), and quartzose (Q). The rationale here is that a component below 10% can
NU
be ignored for nomenclatural purposes.

These 15 fields are sufficiently narrow to be meaningful and at the same time sufficiently wide to
MA

allow samples to be classified by a careful inspection under the microscope, even without the need

of full quantitative analysis. Difficulties may arise for samples in which quartz, feldspar, and lithic
ED

fragments are all present in subequal proportions (i.e., ≥ 30%QFL), in which case a QFL field

may be informally considered. Observations of sand and sandstone specimens with a hand lens can
T
EP

hardly be accurate enough to apply this classification in the field, where more generic terms such as

quartzose, feldspar-bearing, lithic-bearing, feldspar-rich, or lithic-rich may be used.


C

In conformity with the original use by Crook (1960) and Dickinson (1970, 1985), and differently
AC

from Weltje, 2006), the prevalent component goes last, so that a litho-feldspatho-quartzose sand has

more quartz than feldspar, and more feldspar than lithics. The main drawback with this

nomenclature is a certain awkwardness of lengthy labels such as litho-feldspatho-quartzose but, on

the other hand, these composite names add to the precision of petrographic descriptions and are

immediately intelligible in terms of relative composition. This economic scheme does not require

the introduction of new fancy names and gets rid of the use of reference standards to define the

composition of rocks (e.g., type greywacke or type arkose; Huckenholz, 1963), an “archetype”
ACCEPTED MANUSCRIPT
13

doctrine already superseded in many fields of the geosciences, from igneous petrology to

stratigraphy (Dott, 1964 p.626; Ager, 1981 ch.7).

6. LITHIC-RICH AND LITHIC-POOR SAND AND SANDSTONE

When detrital components in sand or sandstone samples are highly diverse and potentially so

PT
numerous, can a classification based on three main framework components only be satisfactorily

complete? The answer to this rhetorical question is that two further specifications are needed to

RI
make the classification system sufficiently informative. For lithic-rich sand and sandstone we may

SC
designate the dominant group (e.g., volcanic, plutonic, metamorphic, or sedimentary) and possibly

specific type (e.g., felsic or mafic, low-grade or high-grade, gneiss or sepentineschist, limestone or
NU
chert) of rock fragments, which are the carriers of most robust provenance information. For lithic-
MA

poor sand and sandstone, instead, we may specify the relative abundance of quartz and feldspar, and

possibly the dominant feldspar type as well (i.e., K or Fk = K-feldspar vs. P or Fp = plagioclase).
ED

Adding independent textural information is also useful, because the relative abundance of quartz,

feldspar, and lithic fragments may be markedly influenced by sample grain size.
T
EP

6.1. The classification of lithic-poor sand and sandstone


C

If lithic fragments are few, then it is essential to retrieve and incorporate in the classificatory label
AC

what quartz and feldspars may reveal. Anorogenic sediments deposited along passive margins and

fed by rivers draining continental interiors typically consist of quartz and feldspar mostly

(continental block provenance of Dickinson, 1985; Sciunnach and Garzanti, 2012). In such cases,

key information is provided by the Q/F ratio, traditionally considered as controlled by the

competition between chemical weathering and unroofing rate in source areas. Feldspar prevails

where granitoid basement is unroofed either very rapidly (tectonic arkose of Folk, 1980; basement

uplift sub-provenance of Dickinson, 1985) or in arid climate (climatic arkose of Folk, 1980),
ACCEPTED MANUSCRIPT
14

whereas quartz dominates in case of intense weathering or extensive recycling of older quartz-rich

sandstone in low-relief shield areas during long periods of tectonic quiescence (craton interior sub-

provenance of Dickinson, 1985).

The necessity to formally distinguish between sand and sandstone with Q/F < 1 and typically P > K

(ideal arkose of Dickinson, 1985) from those with Q/F > 1 and commonly P < K (plagioclase being

widely considered more weatherable than alkali feldspar; Goldich, 1938; Velbel, 1993) has been

PT
felt since Brongniart (1826) formally distinguished between arkose granitoide (Q/F < 1) and arkose

RI
commune (Q/F > 1). Different definitions contemplating diverse conventional boundaries have been

SC
proposed since then. In the classification of Folk (1980 p.127), an arkose contains less than 75%

quartz and more than 18.75% feldspar (Q/F < 4), a subarkose from 75% to 95% quartz and from
NU
2.5% to 25% feldspar (3 < Q/F < 38), and a quartzarenite more than 95% quartz (Q/F > 19).

The scheme proposed here to classify lithic-poor sand and sandstone, particularly fruitful in the
MA

study of modern passive-margin sand generated in the subequatorial climatic belt (Garzanti et al.,

2018a), is based on the Q/F ratio and allows identification of diverse categories required to convey
ED

useful compositional information. Distinguished within the feldspatho-quartzose class are a

feldspar-rich subclass (fFQ; 1 < Q/F < 2), which may be differentiated further into plagioclase-rich
T
EP

if P/K > 2 (pFQ) or K-feldspar-rich if K/P > 2 (kFQ), and a quartz-rich subclass (qFQ; 4 < Q/F <

9). Distinguished within the quartzose class is a pure quartzose subclass (qQ; Q%QFL > 95),
C

corresponding to the quartzarenite of Folk (1980). The complete classification scheme including
AC

these refined subdivisions consists of 18 compositional fields overall (Fig. 3D), allowing us to

capture in one composite name the essential petrographic features of sand and sandstone.

6.2. The description of lithic-rich sand and sandstone

The description of lithic-rich samples poses the opposite problem: an excess of information that can

be hardly compressed in the classificatory label. The challenge was tackled by Folk (1968), who

increased the informative value of his classification by adding auxiliary triangles to the main QFR
ACCEPTED MANUSCRIPT
15

diagram (Fig. 1). His MRF-VRF-SRF ternary plot formally distinguishes among lithic-rich

sandstones containing mainly metamorphic, volcanic, or sedimentary rock fragments – formally

named phyllarenite, volcanic arenite, and sedarenite – and next among sedarenites containing

mainly carbonate, chert, or terrigenous rock fragments – named calclithite, chert-arenite, and shale-

arenite or sandstone-arenite –. Ingersoll and Suczek (1979) introduced the LmLvLs ternary plot, the

Gazzi-Dickinson version of Folk’s MRF-VRF-SRF diagram. Ternary diagrams, however, work

PT
badly wherever detrital components cannot be reduced to three groups sharply, and MRF-VRF-SRF

RI
or LmLvLs plots suffer from at least two major problems. First, the boundary between sedimentary

SC
and metasedimentary grains, as well as between volcanic and metavolcanic grains, is difficult to

define in order to obtain consistent results by different operators (Wolf, 1971; Dickinson, 1985
NU
p.338). Progressively more detailed operational solutions have been suggested in subsequent years

(fig. 4 in Dorsey, 1988; fig. 7 in White et al., 2002; figs. 1 to 4 in Garzanti and Vezzoli, 2003), but
MA

the problem remains, for instance, in the distinction between sparitic limestone or dolostone versus

marble grains. Secondly, grains that carry major provenance information such as ultramafic or fine-
ED

grained plutonic clasts (e.g., granophyre) are not unequivocally assigned. All igneous grains may be

grouped within pole Lv, perhaps including cellular serpentinite (Fig. 4O), whereas foliated
T
EP

serpentineschist (Fig. 4P) may join pole Lm with undetermined ultramafic grains split 50-50%.

Other operational choices, including how to consider chert and carbonate grains, are tackled in
C

section 7, but the main point here is that the spectrum of rock fragments is so wide that any attempt
AC

to handle them with a rigid procedure is doomed to failure in complex natural cases. A flexible

scheme is thus required.

As a simple way to indicate prevalent types of rock fragments (reference to rock fragments rather

than to lithics is preferred here to maximize provenance information), the adjectives plutoniclastic,

metamorphiclastic, or sedimentaclastic can be added to the classificatory label in analogy to the

widely used volcaniclastic. These terms, introduced by Ingersoll (1983 p.1137), were originally

meant to be “genetic and interpretive, as opposed to descriptive”. However, in conformity with the
ACCEPTED MANUSCRIPT
16

leitmotif of this article expressed in section 2, these adjectives may be better used to inform

objectively about the dominant rock-fragment type in the sample without any subjective genetic

implication. Also more specific adjectives such as carbonaticlastic, cherticlastic, basalticlastic,

gneissiclastic, or ultramaficlastic may be used to this goal.

For purely volcaniclastic sediments or sedimentary rocks, names such as rhyodacitic and andesitic

sand (Dickinson, 1985) or rhyolite-arenite and andesite-arenite (Folk, 1980 p.128) have been

PT
proposed. In a similar way, the diverse end-member compositions displayed by modern sand

RI
derived from monolithologic sources within distinct tectono-stratigraphic levels of a lithospheric

SC
section (first-order sampling scale of Ingersoll et al, 1993) were designated as metarhyodacite sand,

slate sand, phyllite sand, schist sand, gneiss sand, kinzigite sand, stronalite sand, granite sand,
NU
gabbro sand, and peridotite sand (Garzanti et al., 2006b). In the general situation of polylithologic

sources and mixed rock-fragment types, none of which dominant, terms such as chert-bearing or
MA

serpentineschist-bearing may be used to highlight the presence of a particularly significant

component. The guideline of this articulated free scheme is to borrow previously proposed criteria
ED

and terms to devise a coherent procedure by which every sample of sand or sandstone can be given

its proper name.


T
EP

7. ROCK FRAGMENTS: THE GOLDMINE OF PROVENANCE INFORMATION


C
AC

Whereas single minerals may derive from diverse rock types, the texture and mineralogy of rock

fragments generally point uniquely to a specific lithology (Fig. 4). Such information is invaluable in

provenance analysis and must not be lost during data collection when using the Gazzi-Dickinson

point-counting method (Gazzi and Zuffa, 1970; Suttner and Basu 1985). As recommended by Gazzi

(1966), this can be done either by using a detailed point-counting sheet that allows the simultaneous

registration of the mineral beneath the cross-hair and of the rock fragment in which the mineral is

located (Fig. 5; table 1 in Zuffa, 1980; table 3 in Zuffa, 1985; table 2 in Fontana et al., 1985), or by
ACCEPTED MANUSCRIPT
17

dedicating a second count specifically to rock-fragment types (Ingersoll and Suczek, 1979 p.1220;

Zuffa, 1987 p.52).

7.1. Classification of rock fragments

The effort to discriminate in great detail among the vast spectrum of rock fragments potentially

encountered in sand and sandstone (Fig. 4) in order to retain the maximum possible level of

PT
information may lead to an unacceptable degree of inhomogeneity in datasets. A robust

RI
classification of rock fragments is thus necessary to ensure comparability among results obtained by

SC
different operators.

To this goal, Dickinson (1970 p.700-701) proposed an operational classification of aphanite to


NU
microgranular lithic grains complemented with textural criteria apt to discriminate consistently

among sedimentary (silty-sandy, argillaceous), volcanic or subvolcanic (vitric, felsitic, microlitic,


MA

lathwork, hypabyssal, volcaniclastic), and metamorphic types (metasedimentary, metavolcanic,

hornfelsic). Building on Graham et al. (1976), Ingersoll and Suczek (1979 p.1220-1221) introduced
ED

additional categories for metamorphic lithics based on their mineralogy and texture (i.e.,
T

polycrystalline mica, quartz-mica tectonite, quartz-mica aggregate, quartz-mica-feldspar aggregate).


EP

Dorsey (1988) subdivided metasedimentary lithics by paying specific attention to their

metamorphic rank into Lm1 (slate, quartzite, slatey siltstone) and Lm2 categories (phyllite-schist,
C

phyllitic quartzite, quartz-mica-albite-aggregate), and included chert within sedimentary grains. As


AC

a further refinement, White et al. (2002) classified metasedimentary lithics according to both

protolith (pelitic vs. felsitic) and rank (Lm1 = slate, metasandstone; Lm2 = phyllite; Lm3 =

micaschist, gneiss). Garzanti et al. (2002a) emphasized the importance of ultramafic detritus,

formally distinguishing between cellular lizardite-serpentinite and foliated antigorite-

serpentineschist grains (Fig. 4 O,P). Garzanti and Vezzoli (2003) proposed a comprehensive

operational classification scheme, illustrated by 24 prototypical grains, that considers four protolith

compositions (pelitic, psammitic or felsic volcanic, carbonatic, mafic volcanic) and six successive
ACCEPTED MANUSCRIPT
18

degrees of increasing metamorphic recrystallization from none to medium-high grade. Overall, a

suite of a hundred rock-fragment types and subtypes including sophisticated categories such as

“amphibole-bearing quartz-feldspar gneiss of rank 4” can be used consistently during routine point-

counting by the Gazzi-Dickinson method (Fig. 5).

This does not mean that identification of rock fragments under the microscope is easy, especially in

strongly diagenetically-altered sandstone samples, where even grain boundaries are hard to

PT
establish. Long-standing problems include difficulties in distinguishing felsitic volcanic rock

RI
fragments from impure unfossiliferous chert (Wolf, 1971), correct identification of pseudomatrix as

SC
groundmass generated by the alteration and deformation of lithic grains (Dickinson, 1970), and the

troublesome distinction between terrigenous extrabasinal versus intrabasinal carbonate and non-
NU
carbonate grains (Zuffa, 1980, 1985; Garzanti, 1991). An analogous thorny issue, represented by the

differentiation of neovolcanic versus paleovolcanic detritus (i.e., grains originated from erosion of
MA

penecontemporaneous vs. notably older volcanic rocks; Zuffa, 1980; Critelli and Ingersoll, 1995),

does not generate problems in the classification of sand and sandstone because coeval and non-
ED

coeval volcanic lithics are all assigned to pole L. This distinction, based on criteria described in

Zuffa (1985, 1987), is nevertheless critical for a correct provenance diagnosis.


T
EP

There is no general consensus on how these problems should be treated operationally, whether we

should surrender to challenging grain identifications or not, and whether lithic grains should be
C

grouped by their origin or by their presumed stability. Idealized concepts such as “stability” or
AC

“maturity” are highly questionable (Garzanti, 2017), because grains may be durable in certain

environments but labile in others. Most detrital minerals including olivine and pyroxene can resist

sediment transport over thousands of kilometers in high-energy fluvial, eolian, and marine

environments (Garzanti et al., 2015a, 2015b), whereas none, including quartz and zircon, can be

considered exempt from chemical vulnerability in hyperhumid equatorial climates or during

diagenesis (Crook, 1968; Cleary and Conolly, 1970; Garzanti et al., 2013a, 2018b). Clarity and

objectiveness, which are the essential requirements of a classification, are lost when subjective
ACCEPTED MANUSCRIPT
19

assumptions are made concerning the relative durability of grain types independently of specific

processes and physico-chemical conditions encountered during erosion, transport, deposition, and

burial.

7.2. Chert and polycrystalline quartz

By giving privilege to stability, Gazzi (1966 p.74) and next Dickinson (1970 p.696) chose to group

PT
in pole Q “the sum of quartz and chalcedony grains of all kinds“. Following Krynine (1948),

initially also Folk (1954) decided to assign chert to pole Q, but next changed his mind in favour of

RI
pole R (Folk, 1980 p.127), as in van Andel (1958), Füchtbauer (1959), Chen (1968), and Okada

SC
(1971) (Table 1). As a conciliatory solution, Dickinson and Suczek (1979) included both

polycrystalline quartz and chert in sub-pole Qp, grouped with monocrystalline quartz in the QtFL
NU
plot, and with lithics in the QmFLt plot (Fig. 1). Recurrence to these two different versions of the
MA

QFL diagram is not straightforward and, because of different field boundaries, may make a

classification equivocal. Considering that the monocrystalline/polycrystalline quartz ratio is highly


ED

dependent on grain size and/or on the convention adopted during point-counting, that chert and

felsic volcanic rock fragments are not invariably distinguishable, and that gradation from pure chert
T

to argillite is also common (Folk, 1980 p.126; Dickinson, 1985 p.336-337), the easiest and most
EP

reasonable escape from the tangle is to group all chert grains to pole L (Zuffa, 1980; Dorsey, 1988)
C

and all quartz grains to pole Q (Fontana et al., 1985). This simple procedure allows using a single
AC

QFL diagram for classification with great gain in simplicity and clarity.

7.3. Carbonate and evaporite grains

Difficulties in identifying the intrabasinal versus extrabasinal origin of carbonate grains, and of

sedimentary grains in general, has long been acknowledged (e.g., Gazzi, 1966 p.76). Zuffa (1980,

1985) provided detailed operational criteria to discriminate intraclasts, bioclasts, ooids, or peloids

generated within the sedimentary basin and coeval with deposition versus limestone or dolostone
ACCEPTED MANUSCRIPT
20

rock fragments derived from erosion of older carbonate rocks outside of the sedimentary basin (Fig.

4A,B). Because such distinction is seldom straightforward, grains that cannot be recognized with

certainty as intrabasinal or extrabasinal are assigned during point-counting to a neutral “limeclast”

category (Wolf, 1965; Blatt et al., 1972 p.460; Zuffa, 1980 p.26). Limeclasts may be next

tentatively reapportioned to either the intrabasinal or extrabasinal group, in the worst case by a 50-

50% split to minimize error.

PT
The problem is even more complicated because non-coeval intrabasinal grains exist as well, such as

RI
those reworked locally from previously deposited layers affected by early cementation or

SC
pedogenesis (e.g., beach rock, eolianites, caliche), or eroded from underlying depositional

sequences during episodes of tectonic uplift or eustatic lowstand (e.g., Garzanti et al., 2003; 2017b).
NU
Such complexities may regard a wide variety of grains, including clay chips, gypsum, glaucony,

chert, or phosphate clasts reworked from mudcracks, soil profiles, sabkhas, or marine hardgrounds
MA

(Garzanti, 1991). In sandstones, crystalline carbonate or evaporite rock fragments may be

misidentified as authigenic cement (Henares et al., 2014), an occurrence of “pseudocement” that


ED

represents a specular case to the occurrence of pseudomatrix.

Problems can be tackled, running the risk of making errors, or skipped. On the one side, Mack
T
EP

(1984 p.218) concluded that “carbonate rock fragments provide important information about

source rock and probably should be included in detrital modes”. On the other side, Dickinson (1985
C

p.336) chose not to recalculate extrabasinal carbonate grains with other lithic fragments “because of
AC

their vastly different geochemical response during weathering and diagenesis, as well as the ease of

confusion with intrabasinal carbonate grains”. Using durability as a criterion for discriminating

certain grain types, however, spoils any attempt to design a descriptive classification by introducing

an unacceptable degree of subjectivity. Carbonate grains prove to be resistant in a vast range of

climatic conditions (Gazzi et al., 1973; Zuffa, 1980; Ingersoll et al., 1987; Garzanti et al., 2002b;

Picard and McBride, 2007), but are easily lost in humid climates in the presence of abundant water

and dissolved carbon dioxide (Schnoor and Stumm, 1986; Stumm and Morgan, 1996 p.188ff; Singh
ACCEPTED MANUSCRIPT
21

and France-Lanord, 2002). In arid climates, even gypsum or anhydrite grains are preserved (Fig.

4C; Henares et al., 2014; Garzanti et al., 2016 p.121). There is no objective rule to determine

whether a grain will survive the diverse physico-chemical environments experienced during one or

more sedimentary cycles, and if we are to neglect carbonate rock fragments because they are

potentially soluble, then we should neglect as well even more mechanically or chemically labile

lithic grains such as shale or gypsum. Few sedimentary lithics will be retained.

PT
Petrographic analysis of sand or sandstone samples poses multiple challenges, but they need to be

RI
faced with the conceptual and technological tools at our disposal, and hopefully won with time one

SC
by one. Whenever we fail, problems should be exposed, otherwise they will never be solved. All

types of extrabasinal lithic fragments, independently of their presumed stability, are considered and
NU
assigned to pole L in the proposed classification scheme.
MA

8. PROVENANCE AND CLASSIFICATION OF MODERN SAND


ED

Contrasting compositions of sand derived from diverse orogenic and anorogenic tectonic domains,

and found in major river systems, deserts, and deep-sea fans, provide a solid objective basis to
T

establish a link between petrographic signature and geology of source terranes. Such an approach,
EP

widely pursued since the last century based on both QFR (Krynine, 1948; Folk, 1968; Potter, 1978)
C

and QFL detrital modes (Dickinson and Suczek, 1979), showed that compositional patterns are not
AC

random, although controlled by several factors superposed to source-rock lithology, including

physical and chemical processes during erosion, transportation, and deposition for modern

sediments (environmental bias) and burial as well for ancient sedimentary rocks (diagenetic bias;

Johnsson, 1993; Weltje, 2012; Basu, 2017). An updated overview is provided in this conclusive

section, which focuses on modern sediments  where source terranes and their geodynamic,

geomorphological, and climatic settings are known and all factors affecting sediment composition

can be verified  and includes unpublished data collected at the Laboratory of Provenance Studies
ACCEPTED MANUSCRIPT
22

(University of Milano-Bicocca) on river sand in North America, northern Europe, northern Asia and

Australia, and on eolian sand from the Kalahari and Sahara sand seas.

8.1. Igneous sources

Major orogenic and anorogenic sources of igneous detritus include magmatic arcs, ophiolitic

complexes, and continental flood basalts. Main characteristic features are the abundance of feldspar

PT
in plutoniclastic sand, of volcanic lithics in volcaniclastic sand, and of mafic to ultramafic lithics in

RI
ophiolite-derived sand.

SC
Magmatic arcs as paradigmatic sources of igneous detritus have been thoroughly investigated by

W.R. Dickinson and co-workers, who documented how composition changes systematically
NU
through time as the plutonic roots of the arc massif are progressively stripped off of their volcanic

cover (Dickinson, 1985; Marsaglia and Ingersoll, 1992; Ingersoll, 2012). Undissected basaltic to
MA

andesitic volcanic provinces shed quartz-poor sand containing mainly microlitic lithic fragments

and plagioclase, thus plotting in the FL field and less commonly in the LF field (e.g., owing to
ED

hydraulic concentration of crystals; Dickinson, 1970 p.705). Sand derived from more felsic
T

rhyodacitic products contains quartz and largely felsitic volcanic lithics and may straddle the qFL
EP

and fQL fields. Wherever the deep-seated batholithic roots of the arc start to be patchily exposed,

quartz and K-feldspar increase at the expense of volcanic lithics and detrital modes may straddle the
C

qLF, qFL, and fQL fields. At an advanced stage of dissection, detritus will plot in the lQF or lFQ
AC

fields and, finally, where only the plutonic arc basement is exposed, in the QF or fFQ fields

(Ingersoll and Eastmond, 2007; Garzanti et al., 2018c).

Volcaniclastic detritus shed from continental flood basalts in rift-related settings is not readily

discriminated from detritus derived from undissected magmatic arcs. Large anorogenic igneous

provinces, however, are typically characterized by bimodal products. The Ethiopia-Yemen Traps,

including both basaltic flows and felsic lavas and ignimbrites (Ayalew et al., 2002; Ukstins et al.,

2002), shed sand straddling the L, FL, qFL, and fQL fields (Garzanti et al., 2001, 2015b). The
ACCEPTED MANUSCRIPT
23

Karoo and Etendeka Traps of southern Africa, the latter including quartz latites at higher

stratigraphic levels (Ewart et al., 2004), supply sand plotting in the L and subordinately FL fields

(Garzanti et al., 2014). The potassic Virunga volcanoes, including silica-saturated latites and

trachytes (Rogers et al., 1998), generate sand containing very little or no quartz and plotting in the

FL and subordinately L fields (Garzanti et al., 2013a). Sand produced in basaltic oceanic islands,

including Iceland, Cape Verde, Tahiti, and Hawaii, also plot in the FL and L fields (Marsaglia,

PT
1993; Dinis et al., 2019).

RI
Because the only quartz that oceanic lithosphere contains is hosted in plagiogranite bodies at the top

SC
of magma chambers (Dilek and Furnes, 2011), ophiolite-derived sand consists almost entirely of

plagioclase and mafic to ultramafic lithic fragments. The ideal petrographic trend in case of
NU
progressive dissection of oceanic lithosphere includes volcaniclastic detritus with dominant

lathwork basaltic to vitric and locally boninitic lithics derived from pillow lavas and sheeted-dike
MA

complexes, followed by plutoniclastic detritus derived from underlying plagiogranite and gabbro,

and finally by ultramafic detritus dominated by cellular serpentinite lithics derived from obducted
ED

mantle rocks (Garzanti et al., 2000, 2002a). In the QFL triangle, such a trend would describe a

counter-clockwise loop starting from the FL field for basalticlastic sand, moving towards the qLF
T
EP

or even F field for theoretically pure gabbro sand, and finally back to the L field for ultramafic sand

(Garzanti et al., 2013b).


C
AC

8.2. Sedimentary and metamorphic sources

Major orogenic and anorogenic sources of sedimentary and metamorphic detritus are fold-thrust

belts, subduction complexes, and crustal sections exposed on rift shoulders or within continental

blocks. Sand composition principally depends on the tectono-stratigraphic level exposed to erosion

in the source area. This was first perceived by Krynine (1948), who envisaged the continental crust

as consisting of sedimentary cover strata generating recycled “quartzite”, underlying metamorphic


ACCEPTED MANUSCRIPT
24

rocks generating lithic-rich “greywacke,” and eventually deeper-seated plutonic rocks generating

feldspar-rich “arkose” (Folk, 1980, p.108).

Beside quartz, however, sedimentary rocks can shed a wide spectrum of lithic fragments, and

detrital modes of sedimentaclastic sand may consequently plot all along the QL leg of the QFL

triangle (e.g., Graham et al., 1976). Depending on their mudrock/sandstone ratio, turbiditic

successions accreted within subduction complexes shed sand plotting in the L and QL fields (e.g.,

PT
Garzanti et al., 1998; Di Giulio et al., 2003), in the QL and LQ fields, or even in the Q and qQ

RI
fields (Garzanti et al., 2013b; Limonta et al., 2015). Wherever parent turbiditic sandstones within

SC
accretionary prisms include common feldspar, daughter sand may plot in the fQL, fLQ, or even in

the lFQ field (Cavazza et al., 1993; Garzanti et al., 2002b; Fontana et al., 2003).
NU
At another extreme, sedimentaclastic detritus genenerated in both orogenic and anorogenic settings

may be represented by pure carbonaticlastic or cherticlastic-carbonaticlastic sand plotting in the L


MA

field. Sand consisting exclusively of limestone and dolostone grains (Fig. 4 A,B) is found in

tropical to middle-latitude regions, including the European Alps, the Apennines, and the Middle
ED

East from the Levant rifted margin to the Zagros Mountains, where arid or semiarid climate

favoured carbonate rather than siliciclastic production along Neotethyan shores through most of the
T
EP

Mesozoic. In the same regions, sedimentaclastic sand characterized or even dominated by chert

grains is derived from pelagic strata deposited originally along distal passive margins of southern
C

Neotethys (e.g., Toscana-Umbria domain in the Apennines, Mamonia complex of south Cyprus,
AC

Hawasina unit in northern Oman; Garzanti et al., 2000, 2002a, 2002b).

Detritus generated from metasedimentary rocks changes progressively in composition with

progressive increase in metamorphic grade. During unroofing of upper crustal levels, quartz

increases from slate sand derived from anchimetamorphic and epimetamorphic metapelites, to

phyllite sand derived from lower-greenschist or blueschist facies schists and calcschists, all plotting

in the QL field. Schist sand derived from upper greenschist-facies schists, paragneisses and marbles

may reach into the fLQ field (Garzanti et al., 2010a). During unroofing of deeper crustal levels,
ACCEPTED MANUSCRIPT
25

feldspar increases in gneiss sand derived from amphibolite-facies granitoid gneisses and plotting in

the lFQ field, in kinzigite sand derived from upper-amphibolite-facies metasediments and straddling

the lFQ and lQF fields, and in stronalite sand derived from lower-crustal granulite-facies

metasediments associated with metagabbros and straddling the lQF and qLF fields (Garzanti et al.,

2006).

As a peculiar case of great provenance value, detritus from mantle rocks that have undergone

PT
eclogite-facies metamorphism during subduction may consist entirely of strongly foliated

RI
antigorite-serpentineschist grains plotting in the L field (Fig. 4P; Garzanti et al., 1998, 2004).

SC
8.3. Large rivers
NU
Transcontinental rivers drain diverse geological domains and are thus typically characterized by

mixed sediment compositions (third-order sampling scale of Ingersoll et al., 1993). Quartz, which is
MA

both widespread in source rocks and durable, commonly predominates. Lithic grains vary in type

and relative abundance, whereas feldspar is generally subordinate but common in case of
ED

widespread igneous source rocks (Fig. 6).


T

Large rivers draining subequatorial cratonic regions in Africa or South America may carry sand
EP

consisting almost exclusively of monocrystalline quartz and thus plotting in the qQ field (Congo,

Okavango, White Nile, Paranà, Uruguay), or containing a few feldspars and plotting in the Q
C

(Niger) or even FQ fields (Zambezi). Sand of the Orinoco River, draining also the Andean retroarc
AC

basin, includes a few sedimentary and low-rank metasedimentary lithics (Q; Limonta et al., 2015).

Feldspatho-quartzose river sand is found at tropical (e.g., Limpopo sand, including volcanic lithics

from anorogenic Karoo basalts; fFQ), middle (Mississippi sand; FQ), and high latitudes (e.g.,

Glomma sand in Norway; qFQ).

Large rivers sourced in orogenic belts in Asia, Europe, or South America carry diverse sedimentary

and metamorphic lithics, and their sand may plot in the fLQ field (Rhein, Po, Danube, Kuban,

Amazon), in the lFQ field (Brahmaputra, Yellow River, Red River, Mekong, Salween; Borges et
ACCEPTED MANUSCRIPT
26

al., 2008; Garzanti et al., 2010b; Nie et al., 2015), or straddle the fLQ and lFQ fields (Indus, Ganga,

Irrawaddy,Yang Tze; Garzanti et al., 2005, 2010b, 2016; Vezzoli et al., 2016). Volcanic lithics

derived from andesitic volcanoes in the Andes and Greater Caucasus are common in sand of the

Amazon and Kuban Rivers, whereas volcanic lithics from the anorogenic Emeishan basalts are

subordinate in Yang Tze sand. Sand of several major rivers in North America and Russia also plot

in the fLQ (Slave, Liard, Colorado) or lFQ fields (St. Lawrence, Lena, Ural), together with bedload

PT
sand of transcontinental rivers such as the Nile and the Orange, which drain continental blocks

RI
including vast anorogenic basaltic provinces and thus carry notable amounts of largely mafic

SC
volcanic lithics. Nile suspended load is instead almost purely volcaniclastic and straddles the qFL

and fQL fields, as sand of the Columbia River that drains both orogenic and anorogenic volcanic
NU
rocks (Whetten et al., 1969).

Quartz-poor river sand plotting in the fQL field is typical of largely undissected orogenic domains
MA

including volcanic regions, such as the Caucasus and the Anatolian-Iranian Plateau (Terek, Rioni,

Kura, Tigris, Euphrates; Vezzoli et al., 2014; Garzanti et al., 2016). European rivers sourced in the
ED

Alpine, Apenninic, or Pyrenean thrust belts and largely draining sedimentary strata may carry

carbonaticlastic sand plotting in the fQL (Rhône, Tevere) or QL (Ebro) fields.


T
EP

None of the considered rivers carries feldspar-dominated sand, but feldspar-poor examples are also

rare. They include the sedimentaclastic sand of the Murray-Darling River in Australia (LQ) and of
C

the Peel River in Canada (QL), as well as the almost purely carbonaticlastic sand of the Karun
AC

River draining the Zagros Mountains in Iran and of several rivers draining the eastern European

Alps (e.g., Piave, Tagliamento; Garzanti et al., 2006b, 2016; Picard and McBride, 2007).

8.4. Deep-sea fans

The composition of deep-sea turbidites chiefly reflects that of their main fluvial feeder system

(Dickinson, 1988; Zuffa et al., 2000). Sand in the huge Indus and Bengal-Nicobar Fans, supplied

from the Himalayas throughout the Neogene, plots in the lFQ field (Ingersoll and Suczek, 1979;
ACCEPTED MANUSCRIPT
27

Suczek and Ingersoll, 1985; Pickering et al., 2018; Andò et al., 2019). Litho-quartzose

metamorphiclastic turbidites characterize the Hellenic Trench (Bartolini et al., 1975), and quartzose

to litho-quartzose and more rarely feldspatho-litho-quartzose compositions characterize Orinoco-

derived turbiditic sandstones exposed on Barbados island (Velbel, 1985; Limonta et al., 2015).

Trench, forearc, and back-arc turbidites all around the Pacific Ocean are chiefly fed from erosion of

magmatic arcs in various stages of dissection (Fig. 7; Dickinson, 1982; Thornburg and Kulm,

PT
1987). Deep-sea sand derived from undissected island arcs or continental arcs mostly consists of

RI
volcanic lithics and plagioclase, and thus plots in the FL field or occasionally in the LF field.

SC
Increasing degrees of dissection are indicated by progressive increase in quartz, K-feldspar,

sedimentary or metamorphic lithics, until only quartz, feldspars, and lithic grains from metamorphic
NU
wallrocks are found where granitoid batholiths are stripped off of their volcanic and sedimentary

cover strata (Marsaglia and Ingersoll, 1992). Sand thus plots in the qFL field or in the adjacent fQL
MA

and qLF fields in the transitional stage (e.g., Yerino and Maynard, 1984; De Rosa et al., 1986;

Marsaglia et al., 1995; Heberer et al., 2010), and straddle the lQF and lFQ fields or even the QF and
ED

fFQ fields when the arc basement becomes the dominant or exclusive source (Bachman and

Leggett, 1982).
T
EP

Quaternary sand in deep-sea fans fed by transcontinental rivers draining anorogenic continental

domains largely consists of quartz and feldspar. Congo Fan sand is quartzose, whereas Nile Cone
C

sand plots in the qFQ field and contains volcanic lithics derived from the Ethiopian continental
AC

flood basalts (Bartolini et al., 1975; Garzanti et al., 2018b).

8.5. Deserts

Dune fields occupy wide areas in arid tropical and subtropical regions. Quartz is commonly

dominant because of its mechanical and chemical durability (Muhs, 2004), and eolian sand plots

mostly in the qQ field where it is largely recycled from older quartzose sandstone, as in the

Kalahari Desert of southern Africa or in the Nafud and Dahna Deserts of Arabia (Fig. 8; Garzanti et
ACCEPTED MANUSCRIPT
28

al., 2013c, 2014). For the same reason, quartz is dominant across the Sahara, although dunes in the

Western Desert of Egypt and in Sinai may contain feldspar and mainly carbonate rock fragments

closer to exposed Mesozoic and Cenozoic carbonates, thus plotting in the qFQ or lFQ fields. Sand

accumulated in coastal deserts along the hyperarid Atlantic coast of southwestern Africa, all largely

fed from the Orange River, includes significant basaltic detritus and plots in the FQ or lFQ fields.

Sand seas of central Asia, such as the Taklamakan and Karakum, largely fed by the Yarkhand and

PT
Amu Darya fluvial systems, commonly include carbonate and other sedimentary or

RI
metasedimentary lithics and may plot in the fQL, fLQ, or lFQ fields (Rittner et al., 2016; Garzanti

SC
et al., 2019).
NU
9. CONCLUSIONS
MA

“The perfect classification of sandstones does not now and never will exist” (Folk, 1980, p.126).

Indeed, the information contained in a sand or sandstone sample is so rich that it could never be
ED

compressed in a short classificatory label. Nature is never trivially repetitive, and the interplay of

physical and chemical processes that generate sand produces an endless suite of products that
T

cannot be captured in a finite set of standard cages. Most classifications proposed in the past have
EP

revealed the ambiguities inevitably implied in the hopeless attempt to encapsulate in one or two

words  and worse when referring to vaguely defined archetypes such as greywacke or arkose  a
C
AC

range of different properties, including mineralogy and texture as well as hydraulic behaviour, and

the presumed durability of detrital grains to mechanical or chemical attack during one or more

sedimentary cycles. This illusion has generated a conceptual tangle that needs to be unraveled.

This paper reviews the somewhat intricate problems involved in the classification of sand and

sandstone, and prescribes a series of norms apt to communicate efficiently the results of

petrographic observations in a scientific article or technical report. For quantitative analysis, the

Gazzi-Dickinson method is recommended in order to get rid of the geometrical effects related to
ACCEPTED MANUSCRIPT
29

breakage of coarse-grained rock fragments into their monomineralic constituents (Ingersoll et al.,

1984; Zuffa, 1985), and thus isolate the component of grain-size-dependent intrasample and

intersample compositional variability associated with provenance and hydraulic sorting in the

depositional environment (Garzanti et al., 2009). It is fundamental to use a detailed point-counting

sheet (e.g., Fig. 5) that allows collection of data on the complete spectrum of rock fragments in the

sample and to recalculate petrographic parameters according to both Gazzi-Dickinson QFL and

PT
traditional QFR methods. All quartz, both monocrystalline and polycrystalline, should be included

RI
in pole Q, and all sedimentary lithics, including chert, extrabasinal carbonate and evaporite, should

SC
be included in pole L (Zuffa, 1980).

The essential information thus obtained can be condensed in a composite adjective based on the
NU
relative abundance of quartz, feldspar, and lithic fragments. According to standard use (Crook,

1960; Dickinson, 1970), the less abundant component goes first, the more abundant last. For lithic-
MA

rich sand and sandstone, information on the prevailing rock fragment type can be communicated by

an additional adjective (Ingersoll, 1983). For lithic-poor sand and sandstone, further formal
ED

subdivisions are proposed, thus reaching a total of 18 compositional fields overall. This scheme

proves to be sufficiently complete to formulate a clear description of any sand or sandstone sample
T
EP

intended to be short and exhaustive, precise and flexible at the same time, as documented with

examples of modern sand in major rivers, eolian dune fields, and deep-sea fans.
C
AC

ACKNOWLEDGMENTS

This article is dedicated to Bob Folk and Bill Dickinson, who taught me far more than the riddles of

sandstone petrography. Ray Ingersoll and Gert Jan Weltje kindly provided masterly reviews,

advice, and extremely useful constructive critical comments. Unpublished petrographic data used in

this study were obtained by Giovanni Vezzoli, Alberto Resentini, and Matteo Sala from samples
ACCEPTED MANUSCRIPT
30

kindly provided by A.Haedke and H.Wittman (world rivers), A. Stone (Kalahari), Y. Najman, Laura

Fielding, J.Roskin, L. Baglioni, R.Bitonte, and D.Roncoroni (Sahara).

Figure 1. Traditional sandstone classifications based on petrographic composition. Note major

differences in nomenclature and definition of end-members and compositional fields. Chert is either

grouped with quartz in Krynine’s, Hubert’s, early Folk’s, and McBride’s QFR plots to emphasize

PT
durability, but with rock fragments in van Andel’s and late Folk’s QFR plots to emphasize

provenance. In the genetic Dickinson’s scheme, based instead on the Gazzi-Dickinson method and

RI
on the inferred correspondence between detrital modes and geodynamic setting, chert is grouped

SC
with quartz in the QtFL plot and with lithic fragments in the QmFLt plot, whereas carbonate rock
NU
fragments are excluded.
MA

Figure 2. Traditional sandstone classifications based on both petrographic composition and texture.

In Pettijohn’s, Gilbert’s, Packham’s, and Crook’s schemes, two different triangular diagrams are
ED

proposed for matrix-rich turbiditic greywackes deposited by high-viscosity fluids distinguished

from cleaner arkose-quartzose arenites deposited by low-density fluids. Diverse tridimensional


T

diagrams were also devised to take into account what Pettijohn (1954 p.360, 363) believed to be the
EP

three factors “of greatest genetic importance in the classification of sandstones”, i.e: “provenance,
C

maturity, and fluidity of the depositing medium”. Adding texture as a fourth component of the
AC

sacred QFR triad, however, involves nomenclatural, graphical, and conceptual confusion (e.g.,

turbidites are not necessarily matrix-rich, pseudomatrix being commonly generated by post-

depositional alteration of labile rock fragments; Okada, 1966; Whetten, 1969; Dickinson, 1970).

Figure 3. The proposed classification of sand and sandstone. A) Straightforward subdivision into

quartzose, feldspathic, and lithic compositions by considering the most abundant QFL component

only. B) In the scheme of Weltje (2006), 6 fields are obtained by considering the relative abundance
ACCEPTED MANUSCRIPT
31

of the two most abundant QFL components. C) In the expanded scheme of Garzanti (2016), 15

fields are obtained by considering all three QFL components provided they exceed 10%QFL. D)

The refined classification of feldspatho-quartzose and quartzose sand and sandstone (Garzanti et al.,

2018a) leads to 18 fields overall. Q = quartzose (qQ = pure quartzose); F = feldspathic; L = lithic;

FQ = feldspatho-quartzose (fFQ = feldspar-rich; qFQ = quartz-rich); QF = quartzo-feldspathic; lF

= litho-feldspathic; FL = feldspatho-lithic; QL = quartzo-lithic; LQ = litho-quartzose; lFQ = litho-

PT
feldspatho-quartzose; lQF = litho-quartzo-feldspathic; qLF = quartzo-litho-feldspathic; qFL =

RI
quartzo-feldspatho- lithic; fQL = feldspatho-quartzo- lithic; fLQ = feldspatho-litho-quartzose.

SC
Figure 4. Rock fragments: the goldmine of provenance information. Sedimentary grains: A, B)
NU
carbonaticlastic sand (Wadi Bih, northern Oman); C) gypsum sand (Azraq dune field, Jordan); D)

chert sand (Romandato creek, southern Italy); E) clast reworked from Oligocene turbidites
MA

(Galathea River, Great Nicobar Island). Metasedimentary grains: F) slate sand (Dazhu River,

Taiwan); G) calcschist from the Tauern Window (Fuschbach River, Austria); H) fibrolitic-
ED

sillimanite schist from the Greater Himalaya (Dordi Khola, Nepal). Volcanic grains derived from:

I) Deccan Trap basalts (Tapti River, India), J) andesites (Rio Grande, Argentina), and K) rhyolites
T
EP

(Lipari Island, Italy). L) Metavolcanic grains from anchimetamorphic Permian rhyolites (Southern

Alps, Italy). Metabasite grains: M) epidote greenschist (Laba River, NW Greater Caucasus) and
C

N) epidote-glaucophane blueschist (Varaita River, W European Alps). Ultramafic grains: O)


AC

cellular serpentinite from obducted ophiolites (Wadi Ham, northern Oman); P) foliated

serpentineschist from subducted metaophiolites (Voltri beach, Italy). All photos taken with crossed

polars; blue bar for scale = 100 m.

Figure 5. Point-counting sheet used routinely for petrographic analysis of modern sand at the

Laboratory of Provenance Studies (Milano-Bicocca University). This scheme allows us to record


ACCEPTED MANUSCRIPT
32

full detailed information on rock-fragment types and to recalculate petrographic parameters

according to both Gazzi-Dickinson QFL and traditional QFR methods.

Figure 6. Sand composition and classification in the world’s largest rivers (mostly own published

and unpublished data). Quartz and sedimentary lithics are generally predominant and feldspar

subordinate (Q/F never < 1) , indicating that in most fluvial systems a large part of the sediment

PT
load is recycled from sedimentary strata. Pure quartzose sand characterizes rivers draining

RI
continental blocks at subequatorial latitudes, where both weathering and recycling are extensive

SC
(White Nile, Congo, Okavango). Lithic-rich sand characterizes rivers draining extensive

sedimentary and volcanic covers in undissected orogenic domains (Euphrates, Tigris, Terek, Kura,
NU
Rhône, Ebro, Tevere). Volcanic lithics are predominant in rivers draining orogenic andesitic arcs

(Amazon), anorogenic basaltic fields (Nile, Orange, Limpopo), or both (Columbia; Whetten et al.,
MA

1969). Metamorphic lithics are predominant in rivers draining either very active orogens (e.g.,

Brahmaputra) or quiescent continental shields (e.g., Glomma). Smaller symbols refer to suspended
ED

load and exemplify two opposite cases: composition is quite similar to bedload in the Irrawaddy

River, but markedly different in the Nile River. Q = quartz; F = feldspar; L = lithics (Lm =
T
EP

metamorphic; Lv = volcanic; Ls = sedimentary).


C

Figure 7. Sand composition and classification in deep-sea fans. Trench, forearc and back-arc
AC

turbidites all around the Pacific Ocean are chiefly fed from erosion of island arcs in the west and

continental arcs in the east. Relative abundance of volcanic lithics, feldspar, and quartz depends on

character of magmatism and degree of dissection of source terranes (Dickinson, 1985). Remnant-

ocean turbidites shed from the Himalayan orogen are rich in quartz and metamorphic or

sedimentary lithics. Deep-sea fans supplied from transcontinental rivers draining continental blocks

are quartz-rich (e.g., Congo Fan) but may include feldspar and volcanic lithics derived from

continental-flood basalts (e.g., Nile Cone). Each point is the mean sand composition from the upper
ACCEPTED MANUSCRIPT
33

part of one deep-sea core. Fields for magmatic-arc provenance are based on data compilations in

Marsaglia and Ingersoll (1992), Ingersoll and Eastmond (2007), and Garzanti et al. (2018c). Data

sources: BL = Bachman and Leggett (1982); Bl = Baltuck et al. (1985); Br = Bartolini et al. (1975);

D = De Rosa et al. (1986); GI = Gergen and Ingersoll (1986); G = Garzanti et al. (2018b) and own

data; IS = Ingersoll and Suczek (1979); M = Marsaglia (2004); Marsaglia et al. (1992, 1995); PH =

Prasad and Hesse (1982); PI = Packer and Ingersoll (1986); S = Stewart (1977, 1978); SI = Suczek

PT
and Ingersoll (1985); YM = Yerino and Maynard (1984). Q = quartz; F = feldspar; L = lithics (Lm

RI
= metamorphic; Lv = volcanic; Ls = sedimentary).

SC
Figure 8. Sand composition and classification in the world’s largest deserts. Quartz is predominant
NU
in the Sahara, Kalahari, and Great Nafud sand seas nourished by extensive recycling of older

quartz-rich sandstones. Limestone lithics may be common in the Rub’ al Khali and other Arabian
MA

deserts and also occur in dunes of Egypt and Sinai. Basaltic lithics and plagioclase characterize

dune fields along the hyperarid coast of Namibia and Angola mostly fed by the Orange River.
ED

Sedimentary and metasedimentary lithics are common in central Asia dune fields, largely supplied

by major rivers draining orogenic belts (Amu Darya, Yarkhand, Yellow River). Q = quartz; F =
T
EP

feldspar; L = lithics (Lm = metamorphic; Lv = volcanic; Ls = sedimentary).


C

Table 1. Criteria used in the classification of sandstone through the last 70 years.
AC

Mineralo Tecto Textu quartz- feldspar- lithic-rich Polycryst Carbo Cla


Article Ye Compos Che Mi
gical nic re, rich rich sands/sto alline nate y
ar ition rt ca
durability settin mud sand/st sand/ston ne quartz grains ma
grai
g conte one e trix
ns
nt
low - kaoli
Krynin 19 X --- X --- orthoquar ark pole Q pole ? po
rank n in
e 48 tzite os Q le
e grayw a R pole
cke F
Folk 19 X X - --- orthoquar ark grayw ack pole R pole --- po ---
54 - tzite os e Q le
- e R
Van 19 X X - --- quartzose ark grayw ack pole Q pole ? --- ---
Andel 58 - sdst. os e R
- e
Pettijo 19 X X - X quartzi ark --- pole R pole --- --- pol
hn 49 - te osi Q e
ACCEPTED MANUSCRIPT
34

- te M

Pettijo 19 X X - X orthoquar ark subgrayw pole R pole --- --- 4th


hn 54- - tzite os acke Q pol
57 - e e
Dappl 19 X X - X quartzose ark grayw ack pole Q pole ? po pol
es et 53 - sdst. os e Q le e
al. - e R R
Willia 19 X X - X quartz arkosic lithic pole Q pole ? --- ---
ms et 53 - Q
arenit arenite/w a areni
al. -
e/ cke te /
w ack w ack
e e

PT
quartz arkosic lithic
Gilber 19 X X - X pole Q pole pole R --- ?
arenit arenit areni
t 54 - Q
- e/ e/ te /
w ack w acke w ack

RI
e e
Bokm 19 X X - X orthoquar ark lithic sdst. pole Q pole ? ? pol

SC
an 55 - tzite os Q e
- e M
Packh 19 X X - X quartzose arkose / labile pole Q pole ? ? pol
am 54 - sdst./ Q e
labile sdst. /
- subgrayw M
grayw ac labile
NU
acke ke grayw ack
e
quartzo feldspath lithic /
Crook 19 X (X) - X pole Q pole --- ? pol
se ic / labile labile
MA

60 - R e
arenit arenite arenit
- M
e e
Huber 19 X X X X orthoquar ark grayw ack pole Q pole --- po ---
t 60 tzite os e Q le
e R
ED

Mc 19 X --- - --- quartzare ark litharenite pole Q pole ? --- ---


Bride 63 - nite os Q
- e
Dott 19 X X - X quartz feldspat lithic Pole Q Pol ? ? 4th
T

64 - arenit hic arenite areni eQ pol


- e
EP

e/ / w acke te /
w ack w ack
e e
Gazzi 19 X X - (X) (protoquar arc (arenite Pole Q pole --- --- ---
C

66 - zite) os litica) Q
- e
AC

Folk 19 X --- ( (X) quartzare ark litharenite Pole Q pole pole R --- ---
68- X nite os R
80 ) e
Dickin 19 X X - --- quartzo feldspathic lithic Pole Q pole --- --- ---
son 70 - Q
se sdst. sdst.
-
sandsto (arkose (grayw a
nes ) cke)
quartzo feldspathic
Dickin 19 X X X --- lithic Poles Qt Pol --- --- ---
se sdst.,
son 85 sandston & Lt es
sandsto arkose e Qt
nes & Lt
Zuffa 19 X --- - --- quartzare ark litharenite Pole Q pole Pole L --- ---
80- - nite os L
85 - e
This 20 X --- - --- quartzo feldspath lithic Pole Q pole Pole L --- ---
one 19 - sand/ston L
se ic
- e
sand/sto sand/sto
ne ne
ACCEPTED MANUSCRIPT
35

REFERENCES
Ager, D.V., 1981. The Nature of the Stratigraphical Record. Halsted Press, New York, 122 p.

Andò, S., Aharonovich, S., Hahn, A., George, S.C., Clift, P.D., Garzanti, E., 2019. Integrating heavy-mineral,
geochemical, and biomarker analyses of Plio-Pleistocene sandy and silty turbidites: a novel approach for
provenance studies (Indus Fan, IODP Expedition 355). Geological Magazine, in press.

PT
Ayalew, D., Barbey, P., Marty, B., Reisberg, L., Yirgu, G., Pik, R., 2002. Source, genesis, and timing of giant
ignimbrite deposits associated with Ethiopian continental flood basalts. Geochimica et Cosmochimica

RI
Acta, 66, 1429-1448.

SC
Bachman, S.B., Leggett, J.K., 1982. Petrology of Middle America Trench and trench slope sands, Guerrero
Margin, Mexico. Initial reports of the Deep Sea Drilling Project, 66, 429-436.
NU
Baltuck, M., von Huene, R., Arnott, R.J., 1985. Sedimentology of the western continental slope of Central
America. Initial reports of the Deep Sea Drilling Project, 84, 921-937.
MA

Bartolini, C., Malesani, P.G., Manetti, P., Wezel, F.C., 1975. Sedimentology and petrology of Quaternary
sediments from the Hellenic trench, Mediterranean ridge and the Nile cone from D.S.D.P., Leg 13, cores.
Sedimentology, 22, 205–236.
ED

Basu, A., 2017. Evolution of siliciclastic provenance inquiries: A critical appraisal. In: Mazumder, R. (Ed.),
Sediment Provenance, Influences on Compositional Change from Source to Sink. Elsevier, ch.2, 5-23,
T

http://dx.doi.org/10.1016/B978-0-12-803386-9.00002-2.
EP

Blatt, H., Middleton, G., Murray, R., 1972. Origin of Sedimentary Rocks, 2nd Edition. Prentice-Hall,
Englewood, New York, 782 p.
C

Bokman, J., 1955, Sandstone classification: Relation to composition and texture. Journal of Sedimentary
AC

Petrology, 25, 201-206.

Boswell, P.G.H., 1960. The term graywacke. Journal of Sedimentary Petrology, 30, 154-157.

Brongniart, A., 1826. L’ arkose, caractères minéralogiques et histoire géognostique de cette roche. Annales des
sciences naturelles, 8, 113-163.

Cavazza, W., Zuffa, G.G., Camporesi, C., Ferretti, C., 1993. Sedimentary recycling in a temperate climate
drainage basin (Senio River, north-central Italy): composition of source rock, soil profiles, and fluvial
deposits. In: Johnsson, M.J., and Basu, A. (Eds.), Processes Controlling the Composition of Clastic
Sediments. Geological Society of America, Special Paper, 284, 247-262.

Chen, P.Y., 1968. A modification of sandstone classification. Journal of Sedimentary Petrology, 38, 54-60.
ACCEPTED MANUSCRIPT
36

Cleary, W.J., Conolly, J.R., 1972. Embayed quartz grains in soils and their significance. Journal of Sedimentary
Petrology, 42, 899-904.

Critelli, S., Ingersoll, R.V., 1995. Interpretation of neovolcanic versus palaeovolcanic sand grains: an example
from Miocene deep‐marine sandstone of the Topanga Group (Southern California). Sedimentology, 42, 83-
804.

Crook, K.A.W., 1960. Classification of arenites. American Journal of Science, 258, 419-428.

Crook, K.A.W., 1968. Weathering and roundness of quartz sand grains. Sedimentology, 11, 171-182.

PT
Crook, K.A.W., 1974. Lithogenesis and geotectonics: The significance of compositional variation in flysch
arenites (graywackes). In: Dott, R.H., Shaver, R.H. (Eds.), Modern and Ancient Geosynclinal

RI
Sedimentation. Society of Economic Paleontologists and Mineralogists, Special Publication 19, 304-310.

Cummins, W.A., 1962. The greywacke problem. Geological Journal, 3, 51-72.

SC
Dapples, E.C., Krumbein, W.C., Sloss, L.L., 1953. Petrographic and lithologic attributes of sandstones. The
Journal of Geology, 61, 291-317.
NU
De Rosa, R., Zuffa, G.G., Taira, A., Leggett, J.K., 1986. Petrography of trench sands from the Nankai Trough,
southwest Japan: Implications for long-distance turbidite transportation. Geological Magazine, 123, 477-
MA

486.

Dickinson, W.R., 1970. Interpreting detrital modes of graywacke and arkose. Journal of Sedimentary Petrology,
40, 695-707.
ED

Dickinson, W.R., 1982, Compositions of sandstones in circum-Pacific subduction complexes and forearc basins.
American Association of Petroleum Geologists Bulletin, 66, 121-137.
T
EP

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, 333-361.
C

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. Springer,
AC

Berlin, 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., Suczek, C.A., 1979. Plate tectonics and sandstone composition. American Association of
Petroleum Geologists Bulletin, 63, 2164-2172.

Di Giulio, A., Ceriani, A., Ghia, E., Zucca, F., 2003. Composition of modern stream sands derived from
sedimentary source rocks in a temperate climate (Northern Apennines, Italy). Sedimentary Geology, 158,
145-161.
ACCEPTED MANUSCRIPT
37

Dilek, Y., Furnes, H., 2011. Ophiolite genesis and global tectonics: Geochemical and tectonic fingerprinting of
ancient oceanic lithosphere. Geological Society of America Bulletin, 123, 387-411.

Dinis, P.A., Pinto, M.C., Rocha, F.T., Garzanti, E., 2019. Detrital record of the denudation of volcanic islands
under sub-tropical climate (Cape Verde). Chemie der Erde, in press.

Doglioni, C., Panza, G.F., 2015. Polarized plate tectonics. Advances in Geophysics, 56, 1-167.

Dorsey, R.J., 1988. Provenance evolution and unroofing history of amodern arc–continent collision: Evidence
from petrography of Plio–Pleistocene sandstones, eastern Taiwan. Journal of Sedimentary Petrology, 58,

PT
208-218.

Dott, R.H., 1964. Wacke, graywacke and matrix - what approach to immature sandstone classification? Journal

RI
of Sedimentary Petrology, 34, 625-632.

Ewart, A., Marsh, J.S., Milner, S.,C., Duncan, A.R., Kamber, B.S., Armstrong, R.A., 2004. Petrology and

SC
geochemistry of Early Cretaceous bimodal continental flood volcanism of the NW Etendeka, Namibia.
Journal of Petrology, 45, 59-138.
NU
Folk, R.L., 1954. The distinction between grain size and mineral composition in sedimentary rock
nomenclature. The Journal of Geology, 62, 344-359.
MA

Folk, R.L., 1966. A review of grain‐size parameters. Sedimentology, 6, 73-93.

Folk, R.L., 1968 and 1980, Petrology of Sedimentary Rocks. Austin (USA), Hemphill Publishing Co., 182 p.
ED

Fontana, D., Zuffa, G.G., Garzanti, E., 1989. The interaction of eustacy and tectonism from provenance studies
of the Eocene Hecho Group Turbidite Complex (South Central Pyrenees, Spain). Basin Research, 2, 223-
237.
T
EP

Fontana, D., Parea, G.C., Bertacchini, M., Bessi, P., 2003. Sand production by chemical and mechanical
weathering of well lithified siliciclastic turbidites of the Northern Apennines (Italy). In: Valloni, R., Basu,
A. (Eds.), Quantitative Provenance Studies in Italy. Memorie Descrittive della Carta Geologica d' Italia, 61,
C

51-60.
AC

Füchtbauer, H., 1959. Zur Nomenklatur der Sedimentgesteine. Erdöl und Kohle, 12, 605-613.

Galloway, W.E., 1974. Deposition and diagenetic alteration of sandstone in northeast Pacific arc-related basins:
Implications for graywacke genesis. Geological Society of America Bulletin, 85, 379-390.

Garzanti, E., 1991. Non-carbonate intrabasinal grains in arenites: Their recognition, significance and
relationship to eustatic cycles and tectonic setting. Journal of Sedimentary Petrology, 61, 959-975.

Garzanti, E., 2016. From static to dynamic provenance analysis—Sedimentary petrology upgraded. Sedimentary
Geology, 336, 3-13.

Garzanti E., 2017. The maturity myth in sedimentology and provenance analysis. Journal of Sedimentary
Research, 87, 353-365
ACCEPTED MANUSCRIPT
38

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., Scutellà, M., Vidimari, C., 1998. Provenance from ophiolites and oceanic allochtons: modern
beach and river sands from Liguria and the Northern Apennines (Italy). Ofioliti, 23/2, 65-82.

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

PT
Aden, Yemen). The Journal of Geology, 109, 277-297.

Garzanti, E., Vezzoli, G., Andò, S., 2002a. Modern sand from obducted ophiolite belts (Oman, U.A.E.). The

RI
Journal of Geology, 110, 371-391.

Garzanti, E., Canclini, S., Moretti Foggia, F., Petrella, N., 2002b. Unraveling magmatic and orogenic

SC
provenances in modern sands: The back-arc side of the Apennine thrust-belt (Italy). Journal of Sedimentary
Research, 72, 2-17.
NU
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,
MA

73, 572-588.

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,
ED

112, 145-164.

Garzanti, E., Andò, S., Vezzoli, G., Ali Abdel Megid, A., El Kammar, A., 2006a. Petrology of Nile River sands
T

(Ethiopia and Sudan): Sediment budgets and erosion patterns. Earth and Planetary Science Letters, 252,
EP

327-341.

Garzanti, E., Andò, S., Vezzoli, G., 2006b. The continental crust as a source of sand (Southern Alps cross-
C

section, Northern Italy). The Journal of Geology 114, 533–554.


AC

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.

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.

Garzanti, E., Resentini, A., Vezzoli, G., Andò, S., Malusà, M.G., Padoan, M., Paparella, P., 2010a. Detrital
fingerprints of fossil continental-subduction Zones (axial belt provenance, European Alps). The Journal of
Geology, 118, 341-362.
ACCEPTED MANUSCRIPT
39

Garzanti, E., Andò, S., France-Lanord, C., Vezzoli, G., Galy, V., Najman, Y., 2010b. Mineralogical and
chemical variability of fluvial sediments. 1. Bedload sand (Ganga–Brahmaputra, Bangladesh). Earth and
Planetary Science Letters, 299, 368-381.

Garzanti, E., Andò, S., Vezzoli, G., Lustrino, M., Boni, M., Vermeesch, P., 2012. Petrology of the Namib sand
sea: Long-distance transport and compositional variability in the wind-displaced Orange Delta. Earth-
Science Reviews, 112, 173-189.

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

PT
Rift. The Journal of Geology, 121, 547-580.

RI
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 Andaman-Nicobar Ridge).

SC
Earth-Science Reviews, 123, 113-132.

Garzanti, E., Vermeesch, P., Andó, S., Vezzoli, G., Valagussa, M., Allen, K., Khadi, K.A., Al-Juboury, I.A.,
NU
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-
MA

margin sand (southern Africa). The Journal of Geology, 122, 17-42.

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
ED

Namibia). Sedimentology, 62, 971-996.

Garzanti, E., Andò, S., Padoan, M., Vezzoli, G., El Kammar, A., 2015b. The modern Nile sediment system:
T

Processes and products. Quaternary Science Reviews, 130, 9-56.


EP

Garzanti, E., Al-Juboury, A.I., Zoleikhaei, Y., Vermeesch, P., Jotheri, J., Akkoca, D.B., Allen, M., Andò, S.,
Limonta, M., Padoan, M., Resentini, A., Rittner, M., Vezzoli, G., 2016. The Euphrates-Tigris-Karun river
C

system: Provenance, recycling and dispersal of quartz-poor foreland-basin sediments in arid climate. Earth-
Science Reviews, 162, 107-128.
AC

Garzanti, E., Radeff, G., Malusà, M., 2017a. Slab breakoff: A critical appraisal of a geological theory as applied
in space and time. Earth-Science Reviews, 177, 303-319.

Garzanti E., Vermeesch, P., Al-Ramadan, K.A., Andò, S., Limonta, M., Rittner, M., Vezzoli, G., 201b7. Tracing
transcontinental sand transport: From Anatolia-Zagros to the Rub' al Khali Sand Sea. Journal of
Sedimentary Research, 87, 1196-1213.

Garzanti, E., Dinis, P., Vermeesch, P., Andò, S., Hahn, A., Huvi, J., Limonta, M., Padoan, M., Resentini, A.,
Rittner, M., Vezzoli, G., 2018a. Dynamic uplift, recycling, and climate control on the petrology of passive-
margin sand (Angola). Sedimentary Geology, 375, 86-104.
ACCEPTED MANUSCRIPT
40

Garzanti, E., Andò, S., Limonta, M., Fielding, L., Najman, Y., 2018b. Diagenetic control on mineralogical
suites in sand, silt, and mud (Cenozoic Nile Delta): Implications for provenance reconstructions. Earth-
Science Reviews, 185, 122-139.

Garzanti, E., Limonta, M., Vezzoli, G., An, W., Wang, J.G., Hu, X.M., 2018c. Petrology and multimineral
fingerprinting of modern sand generated from a dissected magmatic arc (Lhasa River, Tibet). In: Ingersoll,
R.V., Lawton, T.F., Graham, S.A. (Eds.), Tectonics, Sedimentary Basins, and Provenance: A Celebration
of William R. Dickinson’s Career. Geological Society of America, Special Paper 540,
https://doi.org/10.1130/2018.2540(09).

PT
Garzanti, E., Ghassemi M.R., Limonta, M., Resentini, A., 2019. Provenance of Karakum desert sand
(Turkmenistan): Lithic-rich orogenic signature of central Asian dune fields. Rivista Italiana di

RI
Paleontologia e Stratigrafia, in press.

SC
Gazzi, P., 1966. Le arenarie del flysch sopracretaceo dell’ Appennino modenese: Correlazioni con il flysch di
Monghidoro. Mineralogica Petrographica Acta, 12, 69-97.
NU
Gazzi, P., Zuffa, G.G., 1970. Le arenarie paleogeniche dell’Appennino emiliano. Mineralogica et Petrographica
Acta, 16, 97-137.
MA

Gazzi, P., Zuffa, G.G., Gandolfi, G., Paganelli, L., 1973. Provenance and dispersal of the sands of the Adriatic
beaches between the Isonzo and Foglia mouths: Regional setting. Società Geologica Italiana, Memorie, 12,
1-37.
ED

Gergen, L.D., Ingersoll, R.V., 1986. Petrology and provenance of Deep Sea Drilling Project sand and sandstone
from the north Pacific Ocean and the Bering Sea. Sedimentary Geology, 51, 26-56.
T

Gilbert, C.M., 1954. Sedimentary rocks. In: Williams, H., Turner, F.J., Gilbert, C.M., Petrography. San
EP

Francisco, Freeman, 406 p.

Goldich, S.S., 1938. A study in rock weathering. The Journal of Geology, 46, 17-58.
C

Graham, S.A., Ingersoll, R.V., Dickinson, W.R., 1976. Common provenance for lithic grains in Carboniferous
AC

sandstones from Ouachita Mountains and Black Warrior Basin. Journal of Sedimentary Petrology, 46, 620-
632.

Hamilton, W.B., 2011. Plate tectonics began in Neoproterozoic time, and plumes from deep mantle have never
operated. Lithos, 123, 1-20.

Heberer, B., Röser, G., Behrmann, J.H., Rahn, M., Kopf, A., 2010. Holocene sediments from the Southern Chile
Trench: a record of active margin magmatism, tectonics and palaeoseismicity. Journal of the Geological
Society, London, 167, 539-553.

Helmbold, R., van Houten, F.B., 1958. Contribution to the petrography of the Tanner Graywacke. Geological
Society of America Bulletin, 69, 301-342.
ACCEPTED MANUSCRIPT
41

Henares, S., Bloemsma, M.R., Donselaar, M.E., Mijnlieff, H.F., Redjosentono, A.E., Veldkamp, H.G., Weltje,
G.J., 2014. The role of detrital anhydrite in diagenesis of aeolian sandstones (Upper Rotliegend, The
Netherlands): Implications for reservoir-quality prediction. Sedimentary Geology, 314, 60-74.

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, 914-918.

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.

PT
Ingersoll, R.V., 1990. Actualistic sandstone petrofacies: discriminating modern and ancient source rocks.
Geology, 18, 733-736.

RI
Ingersoll, R.V., 2012. Composition of modern sand and Cretaceous sandstone derived from the Sierra Nevada,
California, USA, with implications for Cenozoic and Mesozoic uplift and dissection. Sedimentary Geology,

SC
280, 195-207.

Ingersoll, R.V., Suczek, C.A., 1979. Petrology and provenance of Neogene sand from Nicobar and Bengal Fans,
NU
DSDP Sites 211 and 218. Journal of Sedimentary Petrology, 49, 1217-1228.

Ingersoll, R.V., Eastmond, D.J., 2007. Composition of modern sand from the Sierra Nevada, California, U.S.A.:
MA

implications for actualistic petrofacies of continental-margin magmatic arcs. Journal of Sedimentary


Research, 77, 784-796.

Ingersoll, R.V., Bullard, T.F., Ford, R.L., Grimm, J.P., Pickle, J.D., Sares, S.W., 1984. The effect of grain size
ED

on detrital modes: A test of the Gazzi-Dickinson point-counting method. Journal of Sedimentary Petrology
54, 103-116.
T

Ingersoll, R.V., Cavazza, W., Graham, S.A. and Indiana University graduate field seminar participants, 1987.
EP

Provenance of impure calclithites in the Laramide foreland of southwestern Montana. Journal of


Sedimentary Petrology, 57, 995-1003.
C

Ingersoll, R.V., Kretchmer, A.G., Valles, P.K., 1993. The effect of sampling scale on actualistic sandstone
AC

petrofacies. Sedimentology, 40, 937-953.

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, 1-19.

Kay, M., 1951. North American geosynclines. Geological Society of America, Memoir 48, 143 p.

Klein, G.deV., 1963. Analysis and review of sandstone classifications in the North American geological
literature, 1940-1960. Geological Society of America Bulletin, 74, 555-576.

Korenaga, J., 2013. Initiation and evolution of plate tectonics on Earth: theories and observations. Annual
Review of Earth and Planetary Sciences, 41, 117-151.
ACCEPTED MANUSCRIPT
42

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.

Lasius, G., 1789. Beobachtungen über die Harzgebirge, nebst einem Profilrisse, als ein Beytrag zur
mineralogischen Naturkunde. Hannover, Helwingischen Hofbuchhandlung, 132-152.

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, 27, 696-713.

PT
Mack, G.H., 1984. Exceptions to the relationship between plate tectonics and sandstone composition. Journal of
Sedimentary Petrology, 54, 212-220.

RI
Marsaglia, K.M., 1993. Basaltic island sand provenance. In: Johnsson, M.J., Basu, A., (Eds.), Processes
Controlling the Composition of Clastic Sediments. Geological Society of America. Special Paper 284, 41-

SC
65.

Marsaglia, K.M., 2004. Sandstone detrital modes support Magdalena fan displacement from mouth of Gulf of
NU
California. Geology, 32, 45-48.

Marsaglia, K.M., Ingersoll, R.V., 1992. Compositional trends in arc-related, deep-marine sand and sandstone: a
MA

reassessment of magmatic-arc provenance. Geological Society of America Bulletin, 104, 1637-1649.

Marsaglia, K.M., Ingersoll, R.V., Packer, B.M., 1992. Tectonic evolution of the Japanese Islands as reflected in
modal compositions of Cenozoic forearc and backarc sand and sandstone. Tectonics, 11, 1028-1044.
ED

Marsaglia, K.M., Torrez, X.V., Padilla, I., Rimkus, K.C., 1995. Provenance of Pleistocene and Pliocene sand
and sandstone, ODP Leg 141, Chile Margin. Proceedings of the Ocean Drilling Program, Scientific
T

Results, 141, 133-151.


EP

Mawe, J., 1818. A new descriptive catalogue of minerals, consisting of more varieties than heretofore published
and intended for the use of students who may arrange the specimens they collect. London (UK), Longman,
C

Hurst, Rees, Orme and Brown, 96 p.


AC

McBride, E.F., 1963. A classification of common sandstones. Journal of Sedimentary Petrology, 33, 664-669.

Molinaroli, E., Blom, M., Basu, A., 1991. Methods of provenance determination tested with discriminant
function analysis. Journal of Sedimentary Petrology, 61, 900-908.

Muhs, D.R., 2004. Mineralogical maturity in dunefields of North America, Africa and Australia .
Geomorphology, 59, 247-269.

Murchison, R.I., 1854. Siluria. Murray, London, 523 p.

Okada, H., 1966. Non-greywacke "turbidite" sandstones in the Welsh geosyncline. Sedimentology, 7, 211-232.

Okada, H., 1971. Classification of sandstone: analysis and proposal. The Journal of Geology, 79, 509-525.
ACCEPTED MANUSCRIPT
43

Oriel, S.S., 1949. Definitions of arkose. American Journal of Science, 247, 466-476.

Packer, B.M., Ingersoll, R.V., 1986. Provenance and petrology of Deep Sea Drilling Project sands and
sandstones from the Japan and Mariana forearc and backarc regions. Sedimentary Geology, 51, 5-28.

Packham, G.H., 1954. Sedimentary structures as an important factor in the classification of sandstones.
American Journal of Science, 252, 466-476.

Pettijohn, F.J., 1948. A preface to the classification of the sedimentary rocks. The Journal of Geology, 56, 112-
117.

PT
Pettijohn, F.J., 1949. Sedimentary Rocks. Harper and Brothers, New York, 526 p.

Pettijohn, F.J., 1954. Classification of sandstones. The Journal of Geology, 62, 360-365.

RI
Picard, M.D., McBride, E.F., 2007. Comparison of river and beach sand composition with source rocks,

SC
Dolomite Alps drainage basins, northeastern Italy, In: Arribas, J., Johnsson, M.J., Critelli, S. (Eds.),
Sedimentary Provenance and Petrogenesis: Perspectives from Petrography and Geochemistry. Geological
Society of America, Special Paper 420, 1-12.
NU
Pickering, K.T., Pouderoux, H., Carter, A., Andò, S., Garzanti, E., Limonta, M., Vezzoli, G., Milliken, K.L.,
Chemale, F., Mukoyoshi, H., Kutterolf, S., 2018. Sediment provenance and depositional history of the
MA

Nicobar Fan (Bengal Depositional System) from IODP Expedition 362: detrital zircon geochronology,
apatite thermochronometry, sand petrography and heavy-mineral results. American Geophysical Union
2018, Fall Meeting, T23C-0388.
ED

Potter, P.E., 1978. Petrology and chemistry of modern big river sands. The Journal of Geology, 86, 423-449.

Prasad, S., Hesse, R., 1982. Provenance of detrital sediments from the Middle America Trench transect off
T

Guatemala, Deep Sea Drilling Project Leg 67. Initial reports of the Deep Sea Drilling Project, 67, 507-514.
EP

Rittner, M., Vermeesch, P., Carter, A., Bird, A., Stevens, T., Garzanti, E., Andò, S., Vezzoli, G., Dutt, R., Xu,
Z., Lu, H., 2016. The provenance of Taklamakan desert sand. Earth and Planetary Science Letters, 437,
C

127-137.
AC

Rodgers, J., 1950. The nomenclature and classification of sedimentary rocks. American Journal of Science,
248, 297-311.

Rogers, N.W., James, D., Kelley, S.P., De Mulder, M., 1998. The generation of potassic lavas from the eastern
Virunga Province, Rwanda. Journal of Petrology, 39, 1223-1247.

Schnoor, J.L., Stumm, W., 1986. The role of chemical weathering in the neutralization of acidic deposition.
Swiss Journal of Hydrology, 48, 171-195.

Scholle, P.A., 1979. A Color Illustrated Guide to Constituents, Textures, Cements and Porosities of Sandstones
and Associated Rocks. American Association of Petroleum Geologists, Memoir 28, 201 p.
ACCEPTED MANUSCRIPT
44

Sciunnach, D., Garzanti, E., 2012. Subsidence history of the Tethys Himalaya. Earth-Science Reviews, 25, 179-
198.

Sedgwick, A., Murchison, R.I., 1839. Classification of the older stratified rocks of Devonshire and Cornwall.
The London and Edinburgh Philosophical Magazine and Journal of Science, 14 (89), 241-260.

Singh, S.K., France-Lanord, C., 2002. Tracing the distribution of erosion in the Brahmaputra watershed from
isotopic compositions of stream sediments. Earth and Planetary Science Letters, 202, 645-662.

Sorby, H.C., 1880. On the structure and origin of non-calcareous stratified rocks. Proceedings of the Geological

PT
Society London, 36, 46-92.

Spencer, D.W., 1963. The interpretation of grain size distribution of curves of clastic sediments. Journal of

RI
Sedimentary Petrology, 33, 180-190.

Stewart, R.J., 1977. Neogene turbidite sedimentation in Komandorskiy basin, western Bering Sea. American

SC
Association of Petroleum Geologists Bulletin, 61, 192-206.

Stewart, R.J., 1978. Neogene volcaniclastic sediments from Atka basin, Aleutian Ridge. American Association
NU
of Petroleum Geologists Bulletin, 62, 87-97.Stumm, W., Morgan, J.J., 1996. Aquatic Chemistry, John
Wiley & Sons. Inc., New York.
MA

Suczek, C.A., Ingersoll, R.V., 1985. Petrology and provenance of Cenozoic sand from the Indus Cone and the
Arabian Basin, DSDP Sites 221,222, and 224. Journal of Sedimentary Petrology, 55, 340-346.

Suttner, L.J., Basu, A., 1985. The effect of grain size on detrital modes: a test of the Gazzi-Dickinson point-
ED

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
T

Petrology, 51, 1235-1246.


EP

Ukstins, I.A., Renne, P.R., Wolfenden, E., Baker, J., Ayalew, D., Menzies, M., 2002. Matching conjugate
40
volcanic rifted margins: Ar/39 Ar chrono-stratigraphy of pre-and syn-rift bimodal flood volcanism in
C

Ethiopia and Yemen. Earth and Planetary Science Letters, 198, 289-306.
AC

Thornburg, T.M., Kulm. L.D., 1987. Sedimentation in the Chile trench: petrofacies and provenance. Journal of
Sedimentary Petrology, 57, 55-74.

van Andel. T.H., 1958. Origin and classification of Cretaceous, Paleocene and Eocene sandstones of western
Venezuela. American Association of Petroleum Geologists Bulletin, 42, 734-763.

Velbel, M.A., 1985. Mineralogically mature sandstones in accretionary prisms. Journal of Sedimentary
Petrology, 55, 685-690.

Velbel, M.A., 1993. Constancy of silicate-mineral weathering-rate ratios between natural and experimental
weathering: Implications for hydrologic control of differences in absolute rates. Chemical Geology, 105,
89-99.
ACCEPTED MANUSCRIPT
45

Vezzoli, G., Garzanti, E., Vincent, S.J., Andò, S., Carter, A., Resentini, A., 2014. Tracking sediment provenance
and erosional evolution of the western Greater Caucasus. Earth Surface Processes and Landforms, 39,
1101-1114.

Vezzoli, G., Garzanti, E., Limonta, M., Andó, S., Yang, S., 2016. Erosion patterns in the Changjiang (Yangtze
River) catchment revealed by bulk-sample versus single-mineral provenance budgets. Geomorphology,
261, 177-192.

Yerino, L.N., Maynard, J.B., 1984. Petrography of modern marine sands from the Peru‐Chile Trench and
adjacent areas. Sedimentology, 31, 83-89.

PT
Weltje, G.J., 2004. A quantitative approach to capturing the compositional variability of modern sands.

RI
Sedimentary Geology, 171, 59-77.

Weltje, G.J., 2006. Ternary sandstone composition and provenance: An evaluation of the ‘Dickinson model’. In:

SC
Buccianti, A., Mateu-Figueras, G., Pawlowsky-Glahn, V. (Eds.), Compositional data analysis: From theory
to practice. Geological Society of London, Special Publication 264, 611-627.
NU
Weltje, G.J., 2012. Quantitative models of sediment generation and provenance: state of the art and future
developments. Sedimentary Geology, 280, 4-20.
MA

Weltje, G.J., von Eynatten, H., 2004. Quantitative provenance analysis of sediments: Review and outlook.
Sedimentary Geology, 171, 1-11.

Werner, A.G., 1787. Kurze Klassifikation und Beschreibung der verschiedenen Gebürgsarten. Dresden,
ED

Waltherischen Hofbuchhandlung, 28 p.

Whetten, J.T., 1969. Sediments from the lower Columbia River and origin of graywacke. Science, 152, 1057-
T

1058.
EP

Whetten, J.T., Hawkins, J.W., 1970. Diagenetic origin of graywacke matrix minerals. Sedimentology, 15, 347-
361.
C

Whetten, J.T., Kelley, J.C., Hanson, L.G., 1969. Characteristics of Columbia River sediment and sediment
AC

transport. Journal of Sedimentary Petrology, 39, 1149-1166.

White, N., Pringle, M., Garzanti, E., Bickle, M., Najman, Y., Chapman, H., Friend, P., 2002. Constraints on the
exhumation and erosion of the High Himalayan slab, NW India, from foreland basin deposits. Earth
Planetary Science Letters, 195, 29-44.

Wolf, K.H., 1965. Gradational sedimentary products of calcareous algae. Sedimentology, 5, 1-37.

Wolf, K.H., 1971. Textural and compositional transitional stages between various lithic grain types (with a
comment on “Interpreting detrital modes of graywacke and arkose”). Journal of Sedimentary Petrology, 41,
328-332.
ACCEPTED MANUSCRIPT
46

Zuffa, G.G., 1980. Hybrid arenites: Their composition and classification. Journal of Sedimentary Petrology, 50,
21-29.

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, 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, 39-61.

PT
Zuffa, G.G., Normark, W.R., Serra, F., Brunner, C.A., 2000. Turbidite megabeds in an oceanic rift valley
recording jökulhlaups of late Pleistocene glacial lakes of the western United States. The Journal of

RI
Geology, 108, 253-274.

SC
NU
MA
T ED
C EP
AC
Figure 1
Figure 2
Figure 3
Figure 4
Figure 5
Figure 6
Figure 7
Figure 8

You might also like