+ All Categories
Home > Documents > Origin and timing of siderite cementation in Upper Ordovician glaciogenic sandstones from the Murzuq...

Origin and timing of siderite cementation in Upper Ordovician glaciogenic sandstones from the Murzuq...

Date post: 26-Apr-2023
Category:
Upload: strathclyde
View: 0 times
Download: 0 times
Share this document with a friend
13
Origin and timing of siderite cementation in Upper Ordovician glaciogenic sandstones from the Murzuq basin, SW Libya M.A.K. El-ghali a,b, * , K.G. Tajori b , H. Mansurbeg a , N. Ogle c , R.M. Kalin c a Department of Earth Science, Uppsala University, Villava ¨gen 16, SE 75236 Uppsala, Sweden b Department of Earth Science, Faculty of Science, Al-Fateh University, P.O. Box 13696, Tripoli, Libya c School of Civil Engineering, Environmental Engineering Research Centre, The Queen’s University of Belfast, Stranmillis Road, Belfast BT9 5AG, Northern Ireland Received 15 July 2005; received in revised form 8 February 2006; accepted 10 February 2006 Abstract The origin and timing of siderite cementation have been constrained in relation to depositional facies and sequence stratigraphy of Upper Ordovician glaciogenic sandstones from the Murzuq basin, SW Libya. Optical microscope, backscattered electron imagery, and carbon and oxygen stable isotope analysis have revealed that siderite is of eo- and mesogenetic origin. Eogenetic siderite is Mg-poor with a mean composition of (Fe 91.7 Mg 1.5 Ca 0.3 Mn 6.5 )CO 3 , and occurs in paraglacial, tide-dominated deltaic highstand systems tract (HST) sandstones, in paraglacial, foreshore to shoreface HST sandstones and in postglacial, Gilbert-type deltaic lowstand systems tract (LST) sandstones. This siderite is typically of meteoric water origin that influxed into the LST and HST sandstones during relative sea level fall and basinward shift of the strandline. Mesogenetic siderite, which engulfs and thus postdates quartz overgrowths and illite, is Mg-rich with a mean composition of (Fe 72.2 Mg 21.7 Ca 0.8 Mn 5.3 )CO 3 and occurs in the paraglacial, tide-dominated deltaic HST sandstones, in paraglacial foreshore to shoreface HST sandstones, in glacial, tide-dominated estuarine transgressive systems tract (TST) sandstones, in postglacial, Gilbert-type deltaic LST sandstones, and in postglacial, shoreface TST sandstones. d 18 O V-PDB values of this siderite, which range between K22.6 and K13.8‰, suggest that precipitation has occurred from evolved formation waters with d 18 O values between K14.0 and C1.0‰ and was either meteoric, mixed marine– meteoric and/or marine in origin by assuming postdating quartz overgrowths and illite temperature between 80 and 130 8C. q 2006 Elsevier Ltd. All rights reserved. Keywords: Siderite; Glaciogenic sandstone diagenesis; Sequence stratigraphy; Depositional facies; Upper Ordovician; the Murzuq basin; SW Libya 1. Introduction The origin, elemental and isotopic composition, and distribution patterns of siderite cement in sandstones from a wide variety of depositional environments and diagenetic regimes have been the focus of numerous studies (Matsumoto and Iijima, 1981; Curtis et al., 1986; Mozley, 1989; Pye et al., 1990; Morad et al., 1994; Huggett et al., 2000). Siderite typically precipitates from reducing, non-sulphidic pore waters that have evolved in suboxic, methanogenic geochemical conditions (Garrels and Christ, 1965; Froelich et al., 1979; Berner, 1981; Hem, 1985; Morad, 1998). Siderite chemistry has been used to unravel the origin of pore waters which can be either marine, mixed marine–meteoric or meteoric in compo- sition (Curtis and Coleman, 1986; Bahrig, 1989; Mozley, 1989; Mozley and Wersin, 1992; Baker et al., 1996) and all can be influenced by either transgression or regression events (Morad et al., 2000). Defining the geochemistry and distribution of siderite in a sequence stratigraphic context, which is adopted in this study, allows a better understanding of the parameters that control its chemical composition and formation. The depositional facies and sequence stratigraphic framework of the Upper Ordovician glaciogenic sandstones (i.e. glacial, paraglacial, and post- glacial), outlined by El-ghali (2005), made this study feasible. The diagenetic regimes used in this study are: (i) eodiagenesis (0–2 km depth and at less 70 8C), which includes alterations that have occurred where the pore water chemistry was influenced by surface conditions (depositional waters and climate), (ii) mesodiagenesis (depths over 2 km and at Marine and Petroleum Geology 23 (2006) 459–471 www.elsevier.com/locate/marpetgeo 0264-8172/$ - see front matter q 2006 Elsevier Ltd. All rights reserved. doi:10.1016/j.marpetgeo.2006.02.002 * Corresponding author. Present address: Department of Earth Science, Uppsala University, Villava ¨gen 16, SE 752 36, Uppsala, Sweden. Tel.: C46 18 4712552; fax: C46 18 4712591. E-mail addresses: [email protected], [email protected] (M.A.K. El-ghali).
Transcript

Origin and timing of siderite cementation in Upper Ordovician glaciogenic

sandstones from the Murzuq basin, SW Libya

M.A.K. El-ghali a,b,*, K.G. Tajori b, H. Mansurbeg a, N. Ogle c, R.M. Kalin c

a Department of Earth Science, Uppsala University, Villavagen 16, SE 75236 Uppsala, Swedenb Department of Earth Science, Faculty of Science, Al-Fateh University, P.O. Box 13696, Tripoli, Libya

c School of Civil Engineering, Environmental Engineering Research Centre, The Queen’s University of Belfast, Stranmillis Road, Belfast BT9 5AG, Northern

Ireland

Received 15 July 2005; received in revised form 8 February 2006; accepted 10 February 2006

Abstract

The origin and timing of siderite cementation have been constrained in relation to depositional facies and sequence stratigraphy of Upper

Ordovician glaciogenic sandstones from the Murzuq basin, SW Libya. Optical microscope, backscattered electron imagery, and carbon and

oxygen stable isotope analysis have revealed that siderite is of eo- and mesogenetic origin. Eogenetic siderite is Mg-poor with a mean composition

of (Fe91.7Mg1.5Ca0.3Mn6.5)CO3, and occurs in paraglacial, tide-dominated deltaic highstand systems tract (HST) sandstones, in paraglacial,

foreshore to shoreface HST sandstones and in postglacial, Gilbert-type deltaic lowstand systems tract (LST) sandstones. This siderite is typically

of meteoric water origin that influxed into the LST and HST sandstones during relative sea level fall and basinward shift of the strandline.

Mesogenetic siderite, which engulfs and thus postdates quartz overgrowths and illite, is Mg-rich with a mean composition of

(Fe72.2Mg21.7Ca0.8Mn5.3)CO3 and occurs in the paraglacial, tide-dominated deltaic HST sandstones, in paraglacial foreshore to shoreface HST

sandstones, in glacial, tide-dominated estuarine transgressive systems tract (TST) sandstones, in postglacial, Gilbert-type deltaic LST sandstones,

and in postglacial, shoreface TST sandstones. d18OV-PDB values of this siderite, which range between K22.6 and K13.8‰, suggest that

precipitation has occurred from evolved formation waters with d18O values betweenK14.0 andC1.0‰ and was either meteoric, mixed marine–

meteoric and/or marine in origin by assuming postdating quartz overgrowths and illite temperature between 80 and 130 8C.

q 2006 Elsevier Ltd. All rights reserved.

Keywords: Siderite; Glaciogenic sandstone diagenesis; Sequence stratigraphy; Depositional facies; Upper Ordovician; the Murzuq basin; SW Libya

1. Introduction

The origin, elemental and isotopic composition, and

distribution patterns of siderite cement in sandstones from a

wide variety of depositional environments and diagenetic

regimes have been the focus of numerous studies (Matsumoto

and Iijima, 1981; Curtis et al., 1986; Mozley, 1989; Pye et al.,

1990; Morad et al., 1994; Huggett et al., 2000). Siderite

typically precipitates from reducing, non-sulphidic pore waters

that have evolved in suboxic, methanogenic geochemical

conditions (Garrels and Christ, 1965; Froelich et al., 1979;

0264-8172/$ - see front matter q 2006 Elsevier Ltd. All rights reserved.

doi:10.1016/j.marpetgeo.2006.02.002

* Corresponding author. Present address: Department of Earth Science,

Uppsala University, Villavagen 16, SE 752 36, Uppsala, Sweden. Tel.:C46 18

4712552; fax: C46 18 4712591.

E-mail addresses: [email protected], [email protected]

(M.A.K. El-ghali).

Berner, 1981; Hem, 1985; Morad, 1998). Siderite chemistry

has been used to unravel the origin of pore waters which can be

either marine, mixed marine–meteoric or meteoric in compo-

sition (Curtis and Coleman, 1986; Bahrig, 1989; Mozley, 1989;

Mozley and Wersin, 1992; Baker et al., 1996) and all can be

influenced by either transgression or regression events (Morad

et al., 2000).

Defining the geochemistry and distribution of siderite in a

sequence stratigraphic context, which is adopted in this study,

allows a better understanding of the parameters that control its

chemical composition and formation. The depositional facies

and sequence stratigraphic framework of the Upper Ordovician

glaciogenic sandstones (i.e. glacial, paraglacial, and post-

glacial), outlined by El-ghali (2005), made this study feasible.

The diagenetic regimes used in this study are: (i)

eodiagenesis (0–2 km depth and at less 70 8C), which includes

alterations that have occurred where the pore water chemistry

was influenced by surface conditions (depositional waters and

climate), (ii) mesodiagenesis (depths over 2 km and at

Marine and Petroleum Geology 23 (2006) 459–471

www.elsevier.com/locate/marpetgeo

23˚

28

1709

0 500Scale in Kilometres

Al-Hasawnah Mountain

Al H

aruj A

l-Asw

ad

Murzuq Sand Sea

Awbari Sand Sea

Al Wigh

Tajarhi

Al Qatrun

Zuaylah

Murzuq

Awbari

Sabha

Al Fuqaha

Al Awaynat

Tahalah

Ghat

Anay

Ayn Az Zan

NC 174

Brak

˚˚

˚

F2

B1

Idri

N

22

26

30

10 14 18 22

Mediteranian SeaTunisia

Algeria

NigerChad Sudan

Egypt

TripoliBanghazi

Al Haruj Al-Aswad

Jabal As Sawda

Tibisti

Arch

GhadamisBasin

Jadu Basin

0 600KmScale in

22

26

30

10 14 18 22

MurzuqBasin

Al Qarqaf

Older Granite ,Granodiorite

Upper cambrian

UndifferentiatedOrdovician / Camberian

Ordovician

Eolian,Wadi and Fluvio/eolian deposits

Twon and Villages

Unsellected Oil wells

Sellected Oil wells

Legend

Fig. 1. Location map of the study area in the Murzuq basin, SW Libya. Samples used in this study were collected from wells 1 and 2 and from outcrops in areas A–C,

which refer to the Ghat, Al-Qarqaf and Wadi Anlaline areas, respectively.

M.A.K. El-ghali et al. / Marine and Petroleum Geology 23 (2006) 459–471460

temperatures above 70 8C), which is strongly controlled by

increases in temperature and mediated by evolved formation

waters (Morad et al., 2000).

2. Geological setting

The Murzuq basin is composed of a huge Paleozoic

intracratonic basin that covers most of southwestern Libya

(Fig. 1) with an area in excess of 400,000 km2. The basin

contains up to 4000 m of Paleozoic marine deposits truncated

by Mesozoic to Quaternary continental deposits (e.g. Davidson

et al., 2000; Echikh and Sola, 2000; Hallett, 2002). The studied

stratigraphic interval includes part of the Melaz Shuqran and

Mamuniyat formations (ca. 250 m thick; Fig. 2) that were

deposited during the Upper Ordovician (Hirnancian) glaciation

of Gondwana (e.g. Sutcliffe et al., 2000) when the Murzuq

basin was lying along the continental margin of west

Gondwana, close to the ice cap of the South Pole (Kent and

Van der Voo, 1990; Scotese and Barett, 1990; Smith, 1997;

Davidson et al., 2000). Deposition of the Upper Ordovician

glaciogenic sediments has occurred during glacial, paraglacial,

and postglacial periods (El-ghali, 2005) with alternating cold

and warm climatic conditions (Scotese et al., 1999).

The burial history curve, which was constructed from

analyzing shale velocities, suggested that the Upper

Ordovician glacial and glacial-related deposits in one of the

studied wells (Fig. 1) reached a maximum burial depth (ca.

2.6 km) during the late Jurassic to early Paleocene, which was

300 m deeper than present day (Davidson et al., 2000; Fig. 3).

This maximum burial depth corresponded to a maximum

bottom hole temperature of ca. 130 8C, which was ca. 30 8C

higher than present day (Davidson et al., 2000; Fig. 3).

3. Depositional facies and sequence stratigraphy

The short-lived (ca. 0.5–1 million years) Upper Ordovician

glaciation event (Beuf et al., 1971; Deynoux, 1980; Vaslet,

1990; Blanpied et al., 2000) has resulted in the deposition of

the Melaz Shuqran and Mamuniyat formations in the Murzuq

basin, SW Libya (Fig. 2). Detailed depositional facies and

sequence stratigraphy of these deposits were described by

El-ghali (2005), who recognized three depositional sequences

(Fig. 2). These sequences include: (i) depositional sequence 1,

which corresponds to the entire Melaz Shuqran Formation

(Fig. 2) that was deposited during a period of overall

transgression. Glacial advance, loading of the continental

shelf and subsequent glacial retreat caused sea-levels to rise

and thus transgression to occur (Sutcliffe et al., 2000; El-ghali,

2005). This sequence contains: (a) a transgressive systems tract

(TST) with glacial, shoreface to offshore deposits containing

Isos

tatic

rebo

und

cycl

e

Sequ

ence

Stra

ta

Form

atio

n

Scal

e in

(m

)

Lit

holo

gy Grain sizeSedimentary structure

Facies Codes

F S CDm&

HST

Mel

az S

huqr

an

CS

TST

mfs

SB

SB

TST

LST

SB

Mam

uniy

at

HST

TST

ts

mfs

Tanzuft Formation

LST

ts

II

II

I

0

100

150

200

250

300

50

1 2 3 4 5 6 7 8 9 10 11

Cambrian-Ordovician

Firs

t gla

cial

/par

agla

cial

cyc

le

Seco

nd g

laci

al/p

arag

laci

al c

ycle

G

laci

al c

ycle

s

Murzuq Basin, SW - Libya

CS

I

Loading

pebbles

Bioturbation

Striation

Iron nodules

diamictite (Dm)

conglomerate (C)

sandstone (S)

fine grained (F)

Lithofacies

Legend

planar bedding

parallel lamination

current ripples

massive

wave ripples

trough cross-stratification

hummocky cross-stratification

syn-sedimentary deformationstructures

outsized clastic

Sedimentary structure

1 = clay2 = silt

3 = very fine sand4 = fine sand5 = medium sand6 = coarse sand

fine grained

sand

ston

eco

nglo

mer

ate

7 = very coarse sand

8 = granules

9 = pebbles

10 = cobbles11 = bouldrers

Grain size

LST = Lowstand systems tracts

TST = Transgressive systems tracts

HST = Highstand systems tracts

CS = Condensed section

mfs = Maximum flooding surface

ts = Transgressive surface

SB = Sequence boundary

Sequence stratigraphy

Others

lenticular bedding

flaser bedding

mud drapes

Coarsing and shallowing

fining and deepeing

Fig. 2. Schematic sequence stratigraphic, depositional facies and siderite distribution summary for the Upper Ordovician Melaz Shuqran and Mamunyiat formations

in the Murzuq basin (modified after El-ghali, 2005).

M.A.K. El-ghali et al. / Marine and Petroleum Geology 23 (2006) 459–471 461

5 5

SublithareniteSubarkose

Quartz

Quartzarenite

Litharenite

Feldspatic litharenite

Lithic arkose

Ark

ose

Lithic sub-arkose

25

5 5

Subarkose

25

Quartz

Sublith- arenite

Tem

pera

ture

˚C

Age in millions of years

Dep

th in

met

res

0100200300400500

100

501000

2000

130

0

Fig. 3. Burial-thermal history curve of the Upper Ordovician in one of the

studied wells in the Murzuq basin (modified after Davidson et al., 2000).

M.A.K. El-ghali et al. / Marine and Petroleum Geology 23 (2006) 459–471462

ice-rafted debris, and (b) a highstand systems tract (HST) with

paraglacial, tide-dominated deltaic deposits. The HST sand-

stones were deposited when the sediment supply exceeded the

rate of relative sea level rise during glacial retreat (El-ghali,

2005). (ii) Depositional sequence 2 corresponds to the lower

and middle parts of the Mamuniyat Formation (Fig. 2) and

contains: (a) glacial, incised-valley, fluvial deposits of low-

stand systems tract (LST), (b) Glacial, tide-dominated

estuarine and shoreface to offshore TST deposits, and (c)

Paraglacial, foreshore to shoreface HST deposits. The HST

sandstones were deposited when the sediment supply exceeded

the rate of relative sea level rise during glacial retreat (El-ghali,

2005). (iii) Depositional sequence 3 comprises the upper part

of the Mamuniyat Formation (Fig. 2), which was deposited

during a period of isostatic rebound as a result of glacial retreat

and subsequent sea-level fall and formation of a sequence

boundary, SB (Sutcliffe et al., 2000; El-ghali, 2005). The SB is

covered by postglacial, Gilbert-type deltaic (LST) deposits and

postglacial, upper shoreface (TST) deposits. The transgressive

surface (TS) is marked by the presence of transgressive lag

deposits at the base of the upper shoreface deposits, whereas

the maximum flooding surface (MFS) is recognized by the

presence of a condensed section (CS) on top of the TST

deposits in the outer shelf. The CS signifies distal sediment

starvation, which in association with global sea level rise

during the Silurian, marks the boundary between the Upper

Ordovician deposits and the lower Silurian hot shale in the

Murzuq basin (Fig. 2).

5010 10Feldspar Lithic fragments

Fig. 4. Framework grain (i.e. quartz, feldspars and lithic fragments)

composition of 20 representative Upper Ordovician sandstone samples from

paraglacial, tide-dominated HST sandstones, paraglacial, foreshore-shoreface

HST sandstones, glacial, tide-dominated estuarine TST sandstones, postglacial,

Gilbert-type deltaic LST sandstones, and postglacial, shoreface TST sandstones

plotted on McBride’s (1963) classification. The sandstones are quartzarenite in

composition and show no significant variation between depositional facies and

systems tracts.

4. Sampling and analytical procedures

Sandstone samples were collected from outcrops in which

the studied Upper Ordovician stratigraphic intervals are

exposed along the basin margins and from two wells drilled

in the basin center (Figs. 1 and 2). The collected samples

included: (i) paraglacial, tide-dominated deltaic HST

sandstones, (ii) paraglacial, foreshore-shoreface HST sand-

stones, (iii) glacial, tide-dominated estuarine TST sandstones,

(iv) postglacial, Gilbert-type deltaic LST sandstones, and (v)

postglacial, shoreface TST sandstones. Detailed petrographic

examination was performed on thin sections from selected

samples, which were prepared subsequent to impregnation with

blue epoxy under vacuum. The modal compositions were

obtained from 20 representative samples by counting 300

points in each thin section.

Stable carbon and oxygen isotope analysis was carried out

on 24 siderite-cemented sandstone samples; each sample

contained a certain dominant siderite cement type. Isotope

data is presented in the d notation and was measured relative to

the Vienna PDB (PeeDee Belemnite) standard. Analytical

precision was found to be better than G0.04‰. Chemical

analysis of the siderite cements was performed on 24 carbon-

coated, polished thin sections using a Cameca Camebax BX50

microprobe (EMP), equipped with a backscattered electron

detector (BSE). Operational values were such that accelerating

voltageZ20 kV, beam currentZ10–15 nA, and beam spot

sizeZ1–5 mm. Analytical totals of siderites were normalized to

100% for comparison purposes.

5. Results

5.1. Sandstone composition

The sandstones are dominantly medium- to very coarse-

grained, poorly to well-sorted, quartz arenites with a mean

modal composition of Q97.8F1.3L0.9 (Fig. 4; Table 1). Mono-

crystalline quartz was the dominant detrital constituent (52–

80%; av. 71 vol%) and was more abundant than polycrystalline

quartz (trace-18%; av. 2%). Detrital feldspars (trace to 5%; av.

1%) include K-feldspar and plagioclase. The lithic fragments

(trace-2%; av. 1%) are mainly volcanic, low-grade

Table 1

Modal composition (maximum, minimum, average and standard deviation) of 20 representative sandstone samples from paraglacial, tide-dominated HST sandstones, paraglacial, foreshore-shoreface HST

sandstones, glacial, tide-dominated estuarine TST sandstones, postglacial, Gilbert-type deltaic LST sandstones, and postglacial, shoreface TST sandstones of the Upper Ordovician Melaz Shuqran and Mamunyiat

formations in the studied area

Depositional facies and

systems tract

Postglacial shoreface TST (nZ4) Postglacial Gilbert-type deltaic

LST (nZ5)

Paraglacial shoreface HST (nZ6) Glacial tide-dominated estuarine

TST (nZ4)

Paraglacial tide-dominated delatic

HST (nZ4)

Min Max Mean SD Min Max Mean SD Min Max. Mean SD Min Max Mean SD Min Max Mean SD

Detrital composition

Monocrystalline quartz 70.3 78.7 74.9 3.7 52.0 76.0 68.8 9.8 60.0 80.0 67.0 7.1 65.7 75.0 71.3 4.4 67.0 80.0 74.2 5.7

Polycrystalline quartz 0.0 0.7 0.3 0.3 0.3 18.0 5.7 7.3 0.3 2.3 1.1 0.9 1.0 3.0 1.7 0.9 0.0 2.3 1.0 1.2

K-feldspars 0.0 0.0 0.0 0.0 0.0 2.0 0.8 1.0 0.0 3.0 0.8 1.2 0.3 1.0 0.7 0.3 0.0 0.7 0.3 0.4

Plagioclase 0.0 0.0 0.0 0.0 0.0 0.3 0.1 0.2 0.0 2.3 0.7 1.1 0.0 1.3 0.7 0.8 0.0 1.0 0.5 0.6

Volcanic lithic fragments 0.0 0.0 0.0 0.0 0.0 0.7 0.1 0.3 0.0 1.3 0.4 0.5 0.0 2.0 0.8 1.0 0.3 1.7 1.0 0.6

Plutonic lithic fragments 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.3 0.1 0.1 0.0 0.3 0.1 0.2 0.0 0.0 0.0 0.0

Metamorphic lithic frag-

ments

0.0 0.0 0.0 0.0 0.0 0.7 0.2 0.3 0.0 0.7 0.2 0.3 0.0 0.3 0.1 0.2 0.0 0.0 0.0 0.0

Sedimentary lithic frag-

ments

0.0 0.0 0.0 0.0 0.0 1.3 0.3 0.6 0.0 1.0 0.2 0.4 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

Muscovite 0.0 0.0 0.0 0.0 0.0 1.0 0.3 0.4 0.0 1.0 0.3 0.4 0.0 2.7 1.3 1.1 0.0 2.3 1.4 1.0

Chert 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

Matrix 0.0 0.0 0.0 0.0 0.3 10.0 5.7 4.4 0.0 3.3 0.9 1.4 0.0 4.3 1.4 2.0 0.0 3.3 1.0 1.6

Hevy minerals 0.0 0.0 0.0 0.0 0.0 0.7 0.1 0.3 0.0 1.3 0.4 0.5 0.7 1.0 0.8 0.2 0.0 1.0 0.4 0.4

Diagenetic minerals

Kaolinite 0.0 0.0 0.0 0.0 4.0 7.7 5.8 1.8 1.7 6.3 4.6 1.7 1.0 4.3 2.0 1.6 2.3 5.0 3.5 1.1

Clay coating 0.0 1.3 0.7 0.8 1.3 3.7 2.1 1.0 0.0 2.7 0.9 1.2 0.0 0.0 0.0 0.0 0.0 0.7 0.2 0.3

Illite 0.0 0.0 0.0 0.0 0.0 1.3 0.5 0.7 0.0 0.7 0.2 0.3 0.0 0.0 0.0 0.0 0.0 0.3 0.1 0.2

Chlorite 0.0 0.0 0.0 0.0 0.0 0.3 0.1 0.2 0.0 0.7 0.1 0.3 0.0 5.3 1.3 2.7 0.0 0.0 0.0 0.0

Pyrite 0.0 0.0 0.0 0.0 0.0 3.0 0.7 1.3 0.0 0.0 0.0 0.0 0.0 0.7 0.3 0.3 0.0 3.7 0.9 1.8

Quartz overgrowths 0.0 1.0 0.5 0.6 0.0 8.7 4.7 3.3 0.7 8.7 3.6 3.7 1.3 3.7 2.4 1.0 0.0 11.3 4.8 4.8

Glauconite 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.7 0.2 0.3 0.0 1.0 0.6 0.4 0.0 0.0 0.0 0.0

Albitized feldspars 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 1.7 0.6 0.8 0.0 3.0 0.8 1.5 0.0 0.3 0.1 0.2

Siderite 0.0 3.0 0.8 1.5 0.0 2.3 0.6 1.0 0.0 12.0 3.7 4.5 2.3 14.3 7.1 5.3 0.0 15.0 4.8 6.9

Ankarite 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 11.7 2.9 5.8

Dolomite 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 1.7 0.6 0.7 0.0 0.0 0.0 0.0 0.0 1.0 0.3 0.5

Calcite replacement 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

Calcite pore filling 8.3 14.0 11.0 2.6 0.0 2.3 0.5 1.0 0.0 16.7 3.9 6.6 0.0 3.7 1.7 1.9 0.0 6.0 1.5 3.0

Iron oxide 0.0 20.0 5.0 10.0 0.0 1.3 0.3 0.6 0.0 1.7 0.4 0.7 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

Porosity

Intergranular porosity 0.0 11.7 7.3 5.0 0.0 6.7 2.3 3.0 3.3 15.0 7.2 4.2 3.3 7.7 5.3 2.2 0.0 3.0 1.2 1.5

Intragranular porosity 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.7 0.1 0.3 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

Moldic porosity 0.0 0.0 0.0 0.0 0.0 2.7 0.6 1.2 0.0 6.0 2.2 2.3 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

LST, TST, and HST refer to lowstand, transgressive, and highstand systems tracts, respectively.

M.A.K.El-g

haliet

al./Marin

eandPetro

leum

Geology23(2006)459–471

463

M.A.K. El-ghali et al. / Marine and Petroleum Geology 23 (2006) 459–471464

metamorphic and sedimentary sandstones, with trace amounts

of granitic fragments. Mica content (trace-3%; av. 1%)

includes muscovite and a trace amount of biotite. Heavy

minerals (trace-2%; av. !0.5%) includes zircon, apatite, and

epidote. Trace amounts of glauconite grains (trace-1%; av. !0.5% and z100–300 mm in diameter) are rounded to sub-

rounded in shape and are either fresh or reveal various degree

of oxidation. Matrix (trace-10%; av. 2%) is composed mainly

of mud and silt-sized quartz grains, being more abundant in the

postglacial, Gilbert-type deltaic LST sandstones (av. 5%).

There is no significant variation in detrital composition with

respect to the depositional facies and systems tracts (Table 1).

5.2. Siderite: petrography, elemental and stable isotopic

composition, and distribution within depositional facies and

sequence stratigraphy

Siderite cement in the sandstones occurs in trace to

significant amounts (trace-15 vol%; av. 3 vol%; Table 1).

Siderite exhibits coarse-crystalline and, less commonly,

microcrystalline textural habits and displays considerable

variation in chemical composition as revealed by BSE

imaging. Coarse-crystalline siderite is either chemically

homogenous and/or inhomogeneous and occurs as intergra-

nular cement that fills pores varying widely in size (ca. 100–

600 mm across; Fig. 5A and B). Homogenous coarse-crystal-

line siderite is either Mg-poor or Mg-rich, which henceforth

referred to as Types I and II, respectively (Figs. 5C and D, 6A

Fig. 5. Photomicrographs (Crossed Nichols) showing( (A) Coarse-crystalline si

compactional origin. (B) Coarse-crystalline siderite cement that fills smaller interg

suggests mesogenetic origin. Back scattered electronic image showing: (C) homog

paraglacial, tide-dominated HST sandstones, paraglacial, foreshore-shoreface HST

precipitation from meteoric waters that influxed into the sandstones during relative

coarse-crystalline (Type I) siderite associated with kaolinite (arrow) supports preci

and B; Table 2). Type I siderite, which is Mg-poor with a mean

composition of (Fe91.7Mg1.5Ca0.3Mn6.5)CO3 (Table 2; Fig. 5C

and D), tends to fill large intergranular pores (ca. 250–600 mmacross) in loosely packed framework grains (Fig. 5A) and, in

some cases, is closely associated with kaolinite (Fig. 5D). Type

II siderite, which is Mg-rich with a mean composition of

(Fe72.2Mg21.7Ca0.8Mn5.3)CO3 (Table 2; Fig. 6A), tends to fill

relatively small intergranular pores (ca. 100–200 mm across;

Fig. 5B). Type II siderite engulfs, and thus postdating, Type I

siderite (Fig. 6B), quartz overgrowths (Figs. 5B and 6C), and

illite (Fig. 6D).

Inhomogeneous coarse-crystalline siderite, which hence-

forth referred to as Type III, is characterized by mottled texture

under BSE images (Fig. 6E and F). Type III siderite occurs as

scattered patches, which is characterized by Mg-rich compo-

sition (Fe78.6Mg16.9Ca2.2Mn2.3)CO3 embedded in Mg-poor

composition (Fe93.8Mg1.9Ca0.5Mn3.8)CO3 (Table 2; Fig. 6E

and F), which displays some corrosions. Type III siderite tends

to fill relatively smaller intergranular pores compared with

Type I siderite (ca. 100–200 mm across), and is associated with

kaolinite (Fig. 6E) and engulfs, and thus postdates, Type I

siderite crystals (Fig. 6F).

Microcrystalline siderite (!10 mm), which referred to as

Type IV, is chemically homogenous and is Mg-rich with a

mean composition of (Fe70.5Mg22.1Ca0.9Mn6.5)CO3 (Table 2).

Type IV siderite occurs as tiny crystals that fringe detrital

grains, and/or fills intercrystalline micropores between and

dickite crystals. In some cases, the dickite is displaced entirely.

derite cement that fills large, intergranular pores, suggesting an early, pre-

ranular pores and engulfs, and thus postdates, quartz overgrowths, and hence

eneous Mg-poor coarse-crystalline (Type I) siderite, which fills large pores in

sandstones, and postglacial, Gilbert-type deltaic LST sandstones, indicating

sea level fall and basinward shift of the shoreline. (D) Homogeneous Mg-poor

pitation from meteoric waters.

Fig. 6. Back scattered electronic image showing: (A) homogeneous Mg-rich coarse-crystalline (Type II) siderite in paraglacial, tide-dominated HST sandstones,

paraglacial, foreshore-shoreface HST sandstones, glacial, tide-dominated estuarine TST sandstones, postglacial, Gilbert-type deltaic LST sandstones, and

postglacial, shoreface TST sandstones, which is interpreted to have been precipitated from evolved formation water relative to the contemporaneous Upper

Ordovician meteoric waters. (B) Homogenous Mg-poor (Type I) siderite (black arrow) engulfed by, and thus pre-dating, homogenous Mg-rich (Type II) siderite

(white arrow). SEM image showing: (C) homogeneous Mg-rich coarse-crystalline (Type II) siderite engulfing quartz overgrowths. (D) Homogeneous Mg-rich

coarse-crystalline (Type II) siderite engulfing illite. Back scattered electronic image showing: (E) inhomogeneous mottled (Type III) siderite associated with

kaolinite (arrow). (F) Homogenous Mg-poor (Type I) siderite (black arrow) engulfed by, and thus pre-dating, inhomogeneous mottled (Type III) siderite (white

arrow).

M.A.K. El-ghali et al. / Marine and Petroleum Geology 23 (2006) 459–471 465

Siderite cement occurs in various depositional facies and

systems tracts (Fig. 2), including: (i) paraglacial, tide-

dominated deltaic HST sandstones, (ii) paraglacial, foreshore

to shoreface HST sandstones, (iii) glacial, tide-dominated

deltaic TST sandstones, (iv) postglacial, Gilbert-type deltaic

LST sandstones, and (v) postglacial, shoreface TST sand-

stones. The Type I and the Type III siderite cements are

restricted to the paraglacial, tide-dominated deltaic HST

sandstones, paraglacial, foreshore to shoreface HST sand-

stones, and postglacial, Gilbert-type deltaic LST sandstones

close to the sequence boundary (Fig. 2). The Type II siderite

occurs in the paraglacial, tide-dominated deltaic HST

sandstones, paraglacial, foreshore to shoreface HST sand-

stones, glacial, tide-dominated estuarine TST sandstones,

postglacial, Gilbert-type deltaic LST sandstones, and post-

glacial, shoreface TST sandstones (Fig. 2). The Type IV

siderite occurs in the glacial, shoreface to offshore TST and

postglacial, shoreface TST sandstones and along the maximum

flooding surface (Fig. 2).

The stable isotope composition of siderite cements (Table 2)

exhibit depleted d18OV-PDB values that range between K22.6

and K13.8‰. d13CV-PDB values range between –14.5 and

K6.0‰. d18OV-PDB and d13CV-PDB composition of siderite

cements are positively correlated (rZC0.83; Fig. 7).

Table 2

Chemical (mole%) and carbon (d13CPDB‰) and oxygen (d13CPDB‰) stable isotope composition of siderite

Samples Texture Mg(CO3) Ca(CO3) Mn(CO3) Fe(CO3) d13CV-PDB‰ d18OV-PDB‰

F2-4 Type IV 20.1 0.9 4.2 74.8 K8.8 K19.7

F2-5 Type II 18.4 0.5 1.7 79.4 K14.3 K22.6

F2-7 Type II 19.4 0.1 2.7 77.8 K11.2 K18.0

F2-8-1 p1 Type IV 18.4 0.2 5.4 76.0 K10.2 K18.5

F2-8-2 p1 Type I 3.2 0.0 13.2 83.5

F2-9 Type II 25.6 0.5 6.8 67.1 K9.6 K17.6

F2-14 Type II 18.1 1.2 10.2 70.5 K14.5 K21.4

F2-15 Type II 18.4 0.1 6.2 75.3 K10.6 K19.7

F2 16 Type IV 19.7 1.3 9.1 69.9 K8.5 K14.3

F2 17 Type II 22.1 0.3 5.3 72.3 K8.4 K13.8

F1-21 p1 Type III 0.5 1.0 0.8 97.8 K8.5 K14.3

F1-21 p2 Type III 11.4 0.3 2.8 85.6

F1-21 p3 Type III 4.7 0.2 14.9 80.2

F2-22 Type II 24.9 0.2 8.4 66.5 K8.3 K13.9

F2-33 Type II 23.1 0.5 5.3 71.1 K8.7 K14.2

F2-38 Type II 22.7 1.8 6.6 68.9 K12.6 K20.7

F2-41 Type II 21.3 2.1 8.4 68.2 K8.9 K16.0

F2-45 Type II 25.2 0.6 4.1 70.1 K8.6 K14.2

B1-3 Type II 21.5 0.1 4.2 74.2 K8.9 K14.7

B1-4 Type II 20.3 bdl 8.1 71.6 K9.1 K14.9

B1-5-1 p1 Type II 25.5 0.6 4.6 69.3 K7.1 K14.7

B1-5-1 p2 Type II 22.4 2.4 7.4 67.8

B1-5-1 p3 Type I 1.2 0.3 5.4 93.1

B1-5-2 p1 Type III 0.1 0.3 5.3 94.2

B1-5-2 p2 Type III 19.4 1.6 1.0 78.0

B1-5-3 p1 Type IV 15.1 1.2 2.0 81.7

B1-5-3 p2 Type II 16.3 0.8 2.4 80.5

B1-5-4 p1 Type III 0.3 0.6 1.5 97.6

B1-5-4 p2 Type III 19.4 0.1 7.1 73.4

B1-5-4 p3 Type II 12.3 0.6 3.2 83.9

B1-5-4 p4 Type II 26.8 0.5 5.6 67.1

B1-5-6 p1 Type II 25.9 1.3 6.0 66.8

B1-5-6 p2 Type II 24.2 0.5 6.0 69.4

B1-5-5 p1 Type II 22.1 0.4 2.2 75.3

B1-6 Type II 19.9 1.5 7.4 71.2 K11.1 K14.4

B1-7 Type II 22.4 bdl 3.5 74.1 K8.9 K14.7

B1-8 Type IV 26.2 bdl 3.2 70.6 K8.2 K14.5

B1-11 Type II 18.5 bdl 1.1 80.4 K6.0 K15.3

B1-12-1 p1 Type IV 0.2 0.2 0.4 99.2 K7.6 K15.4

B1-12-1 p2 Type III 13.9 5.6 0.4 80.2

B1-12-2 p1 Type II 24.0 0.5 5.9 69.6

B1-12-3 p1 Type I 0.1 0.5 1.0 98.5

B1-12-4 p1 Type III 5.4 0.6 0.4 93.7

B1-12-4 p2 Type III 20.3 3.4 0.3 76.0

B1-12-5 p1 Type II 22.9 0.3 5.3 71.4

B1-12-6 p1 Type II 22.3 0.3 4.6 72.8

B1-12-6 p2 Type IV 25.7 1.4 6.2 66.7

B1-18 Type IV 26.1 0.9 6.1 66.9 K9.2 K14.4

Three varieties of siderite cements have been recognized according to their elemental composition; including homogeneous, Mg-rich (Types II and IV)

siderite with a mean composition of (Fe72.2Mg21.7Ca0.8Mn5.3)CO3 and (Fe70.5Mg22.1Ca0.9Mn6.5)CO3, homogeneous, Mg-poor (Type I) siderite with a mean

composition of (Fe91.7Mg1.5Ca0.3Mn6.5)CO3, and inhomogeneous mottled (Type III) siderite with Mg-rich [(Fe78.6Mg16.9Mn2.2Ca2.3)CO3] patches embedded in

Mg-poor [(Fe93.8Mg1.9Mn0.5Ca3.8)CO3]. Stable isotopic composition represents the samples dominated by Mg-rich and mottled (Types II–IV) siderite cements

and characterized by d18O from K22.6 to K13.8‰ and d13C from K14.5 to K6.0‰.

M.A.K. El-ghali et al. / Marine and Petroleum Geology 23 (2006) 459–471466

6. Discussion

6.1. Paragenesis and origin of siderite cements

Although it is not possible to determine the precise timing of

the various types of diagenetic siderite cement in relation to

other diagenetic alterations, an overall paragenetic sequence

was established based on the textural relationships and oxygen

isotopic data (Fig. 8). Petrographic observations and elemental

composition of siderite cements suggest that the siderite

formed under various diagenetic conditions.

Type I siderite that fills large intergranular pores in loosely

packed framework grains (Fig. 5A) is believed to have an early,

pre-compactional origin, being formed subsequent to kaolinite,

–24 –22 –20 –18 –16 –14 –12–16

–14

–12

–10

–8

–6

–4

R = +0.83 (n = 24)

d18OV-PDB‰

d13C

V-P

DB

Fig. 7. Cross plot of d13CPDB versus d18OPDB values of siderite cements

showing a positive correlation (rZC0.83), which is attributed to increasing

input of 12C from thermal alteration of organic matter during progressive burial

and increasing temperature.

M.A.K. El-ghali et al. / Marine and Petroleum Geology 23 (2006) 459–471 467

which is typically of eogenetic regime (Meisler et al., 1984;

McAulay et al., 1994; Morad et al., 2000; Ketzer et al., 2003).

Conversely, the precipitation of Type II siderite, which fills

smaller pores and engulfs Type I siderite, quartz overgrowths

and illite (Figs. 5B and 6B–D), is believed to have occurred

during deeper burial. Although the exact timing of Type III

siderite is not entirely clear, it engulfs Type I siderite (Fig. 6F)

and this would suggest a later diagenetic origin. Type IV

siderite, which occurs between and, in some cases, displaces

dickite crystals along the maximum flooding surface suggests

later formation during burial diagenesis.

The elemental composition of siderite is obviously

influenced by the concentration of Fe2C, Mg2C, Mn2C, and

Ca2C ions in the pore waters (Matsumoto and Iijima, 1981;

Curtis and Coleman, 1986; Mozley, 1989). The Mg-poor

nature of the eogenetic, Type I siderite (Table 2) suggests

precipitation from meteoric pore waters (Mozley, 1989).

Conversely, Types II and IV siderite indicates precipitation

from pore waters with high Mg concentrations, in other words

Diagenetic minerals eodiagenesis

kaolin

sidertie

pyrite

quartz overgrowths

iron-oxide

450(my)10-25 °C

kaolinite

Type I

Fig. 8. Paragenetic sequence of the diagenetic siderite and associated diagenetic min

formations based on petrographic observation and oxygen isotopic composition and

according to Morad et al. (2000).

it is either of marine (Mozley, 1989) or from evolved formation

water (Curtis and Coleman, 1986; Morad et al., 1994; Rezaee

and Schulz-Rojah, 1998; Rossi et al., 2001). Types II and IV

siderite either precipitate and engulf quartz overgrowths and

illite (i.e. Type II) or precipitated between dickite crystals (i.e.

Type IV). This suggests a mesodiagenetic origin from evolved

formation waters. Type III siderite, which occurs as Mg-rich

patches embedded in Mg-poor siderite with corrosions,

suggests that this siderite formed via partial dissolution and

replacement of Mg-poor siderite by Mg-rich siderite owing to

an increase of Mg ions in the pore waters during subsequent

burial.

The presence of different siderite types in each analyzed

sandstone sample prevents isotopic analysis of individual

siderite cement types. The majority of bulk samples contain

three dominant siderite cements. Although most isotopic

analyses are bulk analysis, it is dominated by Types II–IV

siderite cement. The results can still be used to explain

conditions of siderite precipitation. In order to calculate the

temperature under which the mesogenetic siderite (Types II-

IV) has precipitated, it is important to understand if the evolved

formation waters were originally marine, mixed marine–

meteoric or meteoric. However, precise knowledge of

d18OS-MOW composition of these pore waters is difficult to

achieve. Nevertheless, using the bulk d18OV-PDB values of

these siderite (K22.6 to K13.8‰), the fractionation equation

of Carothers et al. (1988), and assuming d18OS-MOW values (0

toC2‰; Lundegard and Land, 1986) for the formation waters

evolved from the contemporary Upper Ordovician sea water

(K5 to 0‰; Marshall et al., 1997), siderite would have

precipitated at temperatures ranging from 120 to 315 8C. The

uppermost temperatures are unreasonable for siderite precipi-

tation in the Upper Ordovician sandstones because this would

exceed the maximum burial temperature achieved by the

sediments, which is approximately 130 8C (Davidson et al.,

2000; Fig. 3). Therefore, this method of assuming the d18OS-

MOW of the pore waters in order to calculate formation

temperature leads to inconclusive results.

As petrographic observations show that Types II–IV siderite

postdate quartz overgrowths and illite, it is therefore possible to

deduce the d18OV-PDB values of pore waters by assuming the

mesodiagenesis

290(my) 070 °C 130 °C

dickite

Type II and IVType III

erals in the Upper Ordovician sandstones of the Melaz Shuqran and Mamunyiat

burial history curve. The boundary between eodiagenesis and mesodiagenesis is

–20 –10 0 10 20

0

50

100

150

200T

empr

eatu

re (

°C)

d18Owater (S-MOW) ‰

01020

Fig. 9. Range of temperature and d18OS-MOW values of pore waters calculated

based on oxygen isotope values of siderite. Siderite, which is dominated by

homogenous Mg-rich (Types II and III) and inhomogeneous mottled (Type IV)

type, has d18OV-PDB values between K22.6 and K13.8‰. This siderite is

believed to have precipitated during mesodiagenesis at temperatures of 70 to

130 8C from pore waters with d18OS-MOW from K14.0 to C1.0‰. This is

equivalent to evolved formation waters that were originally meteoric, mixed

marine–meteoric and marine.

M.A.K. El-ghali et al. / Marine and Petroleum Geology 23 (2006) 459–471468

temperature range typical for quartz overgrowths and illite

formation (80–130 8C; McBride, 1989; Worden and Morad,

2000). Precipitation would have occurred from pore waters

with d18OS-MOW values between K14.0 and C1.0‰ (Fig. 9).

This wide range of pore water composition suggests that

siderite precipitation has occurred from pore waters varying in

origin from evolved meteoric and mixed marine–meteoric to

marine waters relative to the contemporary Upper Ordovician

meteoric waters (K17‰; Craig and Gordon, 1965), and marine

waters (K5 to 0‰; Marshall et al., 1997), respectively.

The d13CV-PDB values (K14.5 to K6.0‰) of Types II–IV

siderite are consistent a major source of carbon derived from

the thermal decarboxylation of organic matter, which produces

a strong 12C-enriched carbon (Hudson, 1977; Irwin et al., 1977;

Carothers and Kharaka, 1980; Kharaka et al., 1983; Morad,

1998). The positive correlation between d18OV-PDB and

d13CV-PDB values of siderite suggests an increase of 12C from

the thermal alteration of organic matter during progressive

burial and an increase in temperature (Irwin et al., 1977; Morad

et al., 1990).

6.2. Summary model for the distribution of siderite

Integration of petrographic observations and elemental/

isotopic composition of various types of siderite cement in

sandstones have helped to explain its spatial and temporal

distribution patterns. Variations in the chemical composition of

siderite cements are strongly controlled, primarily, by the

chemistry of pore waters, which, in turn, can be linked to

depositional environments (Mozley, 1989). The restriction of

Mg-poor siderite (i.e. Type I) to the paraglacial, tide-

dominated deltaic HST, paraglacial, foreshore to shoreface

HST, and postglacial, Gilbert-type deltaic LST sandstones

close to the SB as well as its close association with eogenetic

kaolinite are interpreted to indicate precipitation during

eodiagenesis by meteoric waters (Fig. 10). The Mg-poor nature

of siderite (i.e. Type I) corresponds to precipitation from

meteoric waters (Mozley, 1989). Meteoric waters were

presumably fluxed into the LST and HST sandstones as a

consequence of relative sea level fall, which was aided by

isostatic rebound due to glacial retreat and subsequent

prograding, and thus basinward shift of the shoreline (El-ghali,

2005). The glacial, tide-dominated estuarine TST and

postglacial, shoreface TST sandstones, which remained poorly

cemented by eogenetic Mg-poor siderite (i.e. Type I), as well

as the paraglacial, tide-dominated deltaic HST, paraglacial,

foreshore to shoreface HST, and postglacial, Gilbert-type

deltaic LST sandstones have all been subjected to cementation

by mesogenetic homogeneous Mg-rich and inhomogeneous

siderite (i.e. Types II–IV; Fig. 10) derived from evolved

formation waters, which were either originally meteoric, mixed

marine–meteoric or marine in composition.

7. Conclusion

Petrographic characteristics and the elemental/stable iso-

topic compositions of siderite cements from Upper Ordovician,

glaciogenic sandstones have allowed interpretation of the

important factors controlling their precipitation and distri-

bution within depositional facies, sequence stratigraphy and

diagenetic evolution. This study has also explained the impact

of changes in the pore water chemistry on the distribution of

eo- and mesogenetic siderite cement. Important findings of the

study include:

(1) Eogenetic Mg-poor (Type I) siderite, which fills large

pores in loosely packed framework grains and is in close

association with eogenetic kaolinite, occurs in paraglacial,

tide-dominated deltaic highstand systems tract sandstone,

paraglacial, foreshore to shoreface highstand systems tract

sandstone, and postglacial, Gilbert-type deltaic lowstand

systems tract sandstone close to the sequence boundary.

This siderite precipitated during eodiagenesis by influx of

meteoric waters into the lowstand systems tracts during

relative sea level fall and basinward shift of the shoreline

and progradation of the highstand systems tracts.

(2) Homogeneous Mg-rich (Types II and IV) and inhomo-

geneous (Type III) siderite cements dominate in the

glacial, tide-dominated estuarine transgressive systems

tract sandstone and postglacial, shoreface transgressive

systems tract sandstone that remained poorly cemented by

eogenetic Mg-poor siderite. They are also prominent in the

paraglacial, tide-dominated deltaic highstand systems

tract, paraglacial, foreshore to shoreface highstand systems

tract, and postglacial, Gilbert-type deltaic lowstand

systems tract sandstones. This siderite is of mesogenetic

postglacial Gilbert-type deltaic glacial fluvial incised-valley

paraglacial foreshore-shoreface MFS

TS

321

MFS

postglacial shoreface landward basinward

TS

paraglacial tide-dominated deltaic

TS+SBglacial shoreface/offshore

Pre-Upper Ordovician sequences

Silurian hot shale sequence

isostatic rebound

MFS

glacial tide-dominated estuarine and offshore

Eod

iage

nesi

sM

esod

iage

nesi

s

1

1

2

2

2

3

3

highstand systems tract

transgressive systems tract

lowstand systems tract

MFSTSSB

maximum flooding surface

sequence boundarytransgressive surface

Mg-poor (Type I) siderite

Mg-rich (Type II) siderite

iron-oxidekaolintized feldspar

quartz overgrowths

kaolinitized feldspar

QtzQtz

Qtz

Qtz Qtz

QtzQtz

Qtz

Qtz Qtz

QtzQtz

Qtz

Qtz Qtz

Qtz

Qtz

QtzQtz

Qtz

Qtz

Qtz

QtzQtz

Qtz

Qtz

QtzQtz

Qtz

Qtz

Qtz

Qtz

Qtz

QtzQtz

Qtz

Qtz

Qtz

Qtz

porosity

feld

SB

SB

dickitized kaolinite

Mg-rich (Type IV)siderite

mottled (Type III) siderite

Qtz

Qtz

QtzQtz

Qtz

Qtz - detrital quartz grains feld- detrital feldspar grains

dickitized kaolinite

Fig. 10. Schematic model showing the spatial and temporal distribution of diagenetic siderite and associated diagenetic minerals as well as variations in the

diagenetic evolution pathways of the Upper Ordovician sandstones within a sequence stratigraphic framework.

M.A.K. El-ghali et al. / Marine and Petroleum Geology 23 (2006) 459–471 469

origin as it fills relatively small pores as well as envelope

quartz overgrowths and illite.

Acknowledgements

M.A.K. El-ghali and K. Tajori dedicate this work to Dr S.

Lagha, Al-Fateh University, Geology Department, Tripoli,

Libya who passed away while he was doing his job, field

geology. The authors thank Dr S. Morad for reading and

commenting on the manuscript. Thanks also go to the

Petroleum Research Center, Tripoli, Libya, especially to

Professor A. Sbeta, Dr A. Bourima and Dr A. El-Harbi for

supporting the fieldwork. The National Oil Corporation,

especially B. El-Mejrab, E. Hamuni and data section staff,

are acknowledged for giving access to the drill core samples.

M.A.K. El-ghali et al. / Marine and Petroleum Geology 23 (2006) 459–471470

References

Bahrig, B., 1989. Stable isotope composition of siderite as indicator of

paleoenvironmental history of oil shale lakes. Papeogeography, Palepcli-

matology, Paleoecology 70, 139–151.

Baker, J.C., Kassan, J., Hamilton, P.J., 1996. Early diagenetic Siderite as

indicator of depositional environment in the Triassic Rewan Group,

southern Bowen Basin, eastern Australia. Sedimentology 43, 77–88.

Berner, R., 1981. A new geochemical classification of sedimentary

environments. Journal of Sedimentary Petrology 51, 359–365.

Beuf, S., Biju-Duval, B., De Charpal, O., Rognon, P., Gariel, O., Bennacef, A.,

1971. Les gres du Paleozoique inferieur du Sahara, 18. Editions Technip,

Paris (Sci. Tech. Petro1e), 464 pp.

Blanpied, C., Deynoux, M., Ghienne, J.F., Rubino, J.L., 2000. Late Ordovician

glacially related depositional systems of the Gargaf Uplift (Libya) and

comparisons with correlative deposits in the Taoudeni Basin (Mauritania).

In: Sola, M.A., Worsley, D. (Eds.), Symposium on Geological Exploration

in Murzuq Basin. Elsevier, Amsterdam, pp. 485–507.

Carothers, W.W., Kharaka, Y.K., 1980. Stable carbon isotopes of HCO3- in oil-

field waters—implication for the origin of CO2. Geochimica et

Cosmochimica Acta 44, 323–332.

Carothers, W.W., Adami, L.H., Rosenbauer, R.J., 1988. Experimental oxygen

isotope fractionation between siderite–water and phosphoric liberated

CO2Ksiderite. Geochimica et Cosmochimica Acta 52, 2445–2450.

Craig, H., Gordon, L.I., 1965. Deuterium and oxygen 18 variations in the

oceans and marine atmosphere, Proceedings of the Spoleto Conference on

Stable Isotopes in Oceanographic Studies and Paleotemperatures. Labor-

atorio di Geologia Nucleare, Pisa, Italy, pp. 9–13.

Curtis, C.D., Coleman, M.L., 1986. Controls on the precipitation of early

diagenetic calcite, dolomite, and siderite concretions in complex

depositional sequence. In: Gautier, D.L. (Ed.), Roles of Organic Matter

in Sediment Diagenesis, 38. Society of Economic Palaeontologists and

Mineralogists, pp. 23–33 (Special Publication).

Curtis, C.D., Coleman, M.L., Love, L.G., 1986. Pore water evolution during

sediment burial from isotopic and mineral chemistry of calcite, dolomite

and siderite concretions. Geochimica et Cosmochimica Acta 50, 2321–

2334.

Davidson, L., Beswetherick, S., Craig, J., Eales, M., Fisher, A., Himmali, A.,

Jho, J., Mejrab, B., Smart, J., 2000. The structure, stratigraphy and

petroleum geology of the Murzuq Basin, southwest Libya. In: Sola, M.A.,

Worsley, D. (Eds.), Symposium on Geological Exploration in Murzuq

Basin. Elsevier, Amsterdam, pp. 295–320.

Deynoux, M., 1980. Les formation glaciaires du Precambrien terminal et de la

fin de l’Ordovicien en Afrique de l’ouest, Deux exemples de glaciation

d’inlandsis sur une plate-forme stable, 17. Trav. Lab. Sci. Terre, Univ. St.

Jerome, Marseille, B., 554 pp.

Echikh, J., Sola, M.A., 2000. Geology and hydrocarbon occurrences in the

Murzuq basin, SW Libya. In: Sola, M.A., Worsley, D. (Eds.), Symposium

on Geological Exploration in Murzuq Basin. Elsevier, Amsterdam,

pp. 175–222.

El-ghali, M.A.K., 2005. Depositional environments and sequence stratigraphy

of paralic glacial, paraglacial and postglacial upper Ordovician deposits in

the Murzuq Basin, SW Libya. Sedimentary Geology 177, 145–173.

Froelich, P.N., Klinkhammer, G.P., Bender, M.L., Luedtke, N.A., Heath, G.R.,

Cullen, D., Dauphin, P., Hammond, D., Hartman, B., Maynard, V., 1979.

Early oxidation of organic matter in pelagic sediments of the eastern

equatorial Atlantic: suboxic diagenesis. Geochimica et Cosmochimica Acta

43, 1075–1090.

Garrels, R.M., Christ, C.L., 1965. Mineral, solutions and Equilibria. Harper and

Row, New York, 450pp.

Hallett, D., 2002. Petroleum geology of Libya. Elsevier science B.V. pp. 503.

Hem, J.D., 1985. Study and interpretation of the chemical characteristics of

natural waters. US Geological Survey Water Supply Paper No. 2254,

263pp.

Hudson, J.D., 1977. Concreations, isotopes, and the diagenetic histroy of the

Oxford Clay (Jurassic) of central England. Sedimentology 25, 339–370.

Huggett, J., Dennis, P., Gale, A., 2000. Geochemistry of early siderite cements

from the Eocene succession of Whitecliff Bay, Hampshire Basin, UK.

Journal of Sedimentary Research 70, 1107–1117.

Irwin, H., Curtis, C., Coleman, M., 1977. Isotopic evidence for source of

diagenetic carbonates formed during burial of organic rich sediments.

Nature 269, 209–2133.

Kent, D.V., Van der Voo, R., 1990. Palaeozoic paleogeography from

palaeomagnetism of Atlantic bordering continents. In: McKerrow, W.S.,

Scotese, C.R. (Eds.), Palaeozoic Paleogeography and Biogeography, 12.

Geological Society Memoir, pp. 49–56.

Ketzer, J.M., Holz, M., Morad, S., Al-Aasm, S., 2003. Sequence stratigraphic

distribution of diagenetic alterations in coal bearing, paralic sandstones:

evidence from the Rio Bonito Formation (early Permian), southern Brazil.

Sedimentology 50, 855–877.

Kharaka, Y.K., Carothers, W.W., Rosenbauer, R.J., 1983. Thermal decarbox-

ylation of acetic acid: implication for origin of natural gas. Geochimica

Cosmochimica Acta 47, 397–402.

Lundegard, P.D., Land, L.S., 1986. Carbon dioxide and organic acids: their role

in porosity enhancement and cementation, Palaeogene of the Texas Gulf

Coast. In: Gautier, D.L. (Ed.), Roles of Organic Matter in Sediment

Diagenesis, 38. Society of Economic palaeontologists and Mineralogists,

pp. 129–146 (Special Publication).

Marshall, J.D., Brenchley, P.J., Mason, P., Wolff, G.A., Astini, R.A., Hints, L.,

Meidla, T., 1997. Global carbone isotopic events associated with mass

extinction and glaciation in the late Ordovician. Palaeogeography,

Palaeoclimatology, Palaeoecology 132, 195–210.

Matsumoto, R., Iijima, A., 1981. Origin and diagenetic evolution of Ca–

Mg–Fe carbonates in some coalfields of Japan. Sedimentology 28,

239–259.

McAulay, G.E., Burley, S.D., Fallick, A.E., Kusznir, N.J., 1994. Palaeohy-

drodynamic fluid flow regimes during diagenesis of the Brent Group in the

Hutton-NW Hutton reservior, constraints from oxygen isotope studies of

authigenic kaolin and reserves flexural modeling. Clay Minerals 29, 609–

629.

McBride, E.F., 1989. Quartz cementation in sandstones—a review. Earth

Science Reviews 26, 69–112.

Meisler, H., Leahy, P.P., Knobel, L.L., 1984. Effect of eustatic sea-level

changes on saltwater–freshwater in the north Atlantic coast plain. United

State Geological Survey (Water-Supply Paper, 2255), 27pp.

Morad, S., 1998. Carbonate cementation in sandstones; distribution patterns

and geochemical evolution. In: Morad, S. (Ed.), Carbonate Cementation in

Sandstones, 26. International Association of Sedimentologist, pp. 1–26

(Special Publication).

Morad, S., Ben Ismail, H., De Ros, L.F., Al-Aasm, I.S., Serrihin, N.E., 1994.

Diagenesis and formation waters chemistry of Triassic reservior sandstones

from southern Tunisia. Sedimentology 41, 1253–1272.

Morad, S., Ketzer, J.M., De Ros, F., 2000. Spatial and temporal distribution of

diagenetic alterations in siliciclastic rocks: implications for mass transfer in

sedimentary basins. Sedimentology 47 (Suppl. 1), 95–120.

Mozley, P.S., 1989. Relationship between depositional environment and the

elemental composition of early digenetic siderite. Geology 17, 704–706.

Mozley, P.S., Wersin, P., 1992. Isotopic composition of siderite as indicator of

depositional environment. Geology 20, 817–820.

Pye, K., Dickson, J.A.D., Schiavon, N., Coleman, M.L., Cox, M., 1990.

Formation of siderite-Mg-calcite-iron sulphide concretions in intertidal

marsh and sandflat sediments, north Norfolk, England. Sedimentology 37,

325–343.

Rezaee, M.R., Schulz-Rojah, J.P., 1998. Application of quantitative back-

scattered electron image analysis in isotope interpretation of siderite

cement: Tirrawarra Sandstone, Cooper basin, Australia. In: Morad, S. (Ed.),

Carbonate Cementation in Sandstones, 26. International Association of

Sedimentologiest, pp. 461–481 (Special Publication).

Rossi, C., Rafaela, M., Ramseyer, K., Permanyer, A., 2001. Facies-related

diagenesis and multiphase sderite cementation and dissolution in the

reservior sandstones of the Khataba formation, Egypts western desert.

Journal of Sedimentary Researsh 71, 459–472.

M.A.K. El-ghali et al. / Marine and Petroleum Geology 23 (2006) 459–471 471

Scotese, C.R., Barett, S.F., 1990. Gondwana’s movement over the south Pole

during the Paleozoic: evidence from lithological indicator of climate. In:

McKerrow, W.S., Scotese, C.R. (Eds.), Paleozoic Paleogeography and

Biogeography, 12. Geological Society of London Memoir, pp. 75–85.

Scotese, C.R., Boucot, A.J.,McKerrow,W.S., 1999.Gondwana paleogeography

and palaeoclimatology. Journal of African Earth Science 28, 99–114.

Smith, A.G., 1997. Estimates of the Earth’s spin (geographic) axis relative to

Gondwana from glacial sediments and palaeomagnetism. Earth Science

Review 41, 161–179.

Sutcliffe, O.E., Dowdeswell, J.A., Whittington, R.J., Theron, J.N., Craig, J.,

2000. Calibrating the late Ordovician glaciation and mass extinction by the

eccentricity cycles of the Earth’s orbit. Geology 18, 967–970.

Vaslet, D., 1990. Upper Ordovician glacial deposits in Saudi Arabia. Episodes

13, 147–161.

Worden, R.H., Morad, S., 2000. Quartz cement in oil field sandstones: a review

of the critical problems. In: Worden, R.H., Morad, S. (Eds.), Quartz

Cementation in Sandstones, 29. International Association of Sedimentol-

ogists, pp. 1–20 (Special Publication).


Recommended