You are on page 1of 16

Marine and Petroleum Geology 48 (2013) 31e46

Contents lists available at ScienceDirect

Marine and Petroleum Geology


journal homepage: www.elsevier.com/locate/marpetgeo

Organic geochemical and petrographic characteristics of Tertiary


coals in the northwest Sarawak, Malaysia: Implications for
palaeoenvironmental conditions and hydrocarbon generation
potential
Mohammed Hail Hakimi a, *, Wan Hasiah Abdullah b, Say-Gee Sia b, Yousif M. Makeen b
a
b

Geology Department, Faculty of Applied Science, Taiz University, 6803 Taiz, Yemen
Department of Geology, University of Malaya, 50603 Kuala Lumpur, Malaysia

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 15 May 2013
Received in revised form
12 July 2013
Accepted 13 July 2013
Available online 24 July 2013

Tertiary coals from Mukah and Balingian coalelds in the northwest Sarawak, Malaysia were investigated
to evaluate their regional rank variation, petroleum generative potential and to reconstruct the palaeoenvironment conditions during peat accumulation. The Tertiary coals are characterized by high total
organic carbon contents (TOC) and yield of bitumen extraction ranging from 31.3 to 55.9 wt. % and
25,724.9-92,143.7 ppm, respectively, meet the standard as a good source rock potential. The Mukah and
Balingian coals were generally plotted in an area of Type III kerogen and mixed Type II/III kerogens with
HI values between 90 and 289 mg HC/g TOC, whereby the coals were derived from plant materials of
terrigenous origin. This shows that the Balingian coals are dominated by Type III terrigenous kerogen
while Mukah coals are dominated by Type II/III kerogens, and are thus considered to be generate mainly
gas-prone and limited oil-prone. This is also supported by macerals composition and open system pyrolysis gas chromatography (Py-GC). The Mukah and Balingian coals are thermally immature in rang
from lignite to sub-bituminous C rank, possessing huminite reectance in the range of 0.26%e0.39%. This
immaturity has a considerable inuence on the proximate analysis, particularly on relatively high
moisture and volatile matter contents and relatively low xed carbon content.
Petrographically, it was observed that the Mukah and Balingian coals are dominated by huminite, with
low to high amounts of liptinite and relatively low amounts of inertinite, pointing to predominantly
anaerobic deposition conditions in the paleomires, with limited thermal and oxidative tissue destruction.
The palaeoenvironment conditions of the coals are generally interpreted as a lower deltaic plain wet
peat-swamp depositional setting, which are generally characterised by low TPI and high GI values, and
are plotted on the marsh eld of the Diessels diagram. This is usually consistent with generate relatively
high ash yield as is the case of the Mukah and Balingian coals.
2013 Elsevier Ltd. All rights reserved.

Keywords:
Tertiary coaleld
Petrology
Coal rank
Petroleum potential
Malaysia

1. Introduction
Coals in Malaysia are present in the Tertiary basins in all three
geographical provinces, viz. Sarawak, Sabah and Peninsular
Malaysia (Fig. 1a). However, most of the coal resources are located
in the states of Sabah and Sarawak (EIA, 2012). As at the end of 2011,
total coal resources in Malaysia amounted at 1819 Mt, of which
1468 Mt or 80.7% were located in Sarawak, 334 Mt or 18.3% in
Sabah, and the remaining 1% in Peninsular Malaysia (EIA, 2012).

* Corresponding author. Geology Department, Faculty of Applied Science,


Taiz University, 6803 Taiz, Yemen. Tel.: 967 73 6517422.
E-mail address: ibnalhakimi@yahoo.com (M.H. Hakimi).
0264-8172/$ e see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.marpetgeo.2013.07.009

Most of the coals are thermal coal and ranging from subbituminous to anthracite; nonetheless coals with coking properties exist in the Bintulu, Silantek, Slimponpon and Maliau coalelds
(Fig. 1a). Several studies concerning sedimentology, geochemical
and organic petrographic characteristics have been performed on
Tertiary coals in the Sarawak and Sabah Basins (e.g., Wan Hasiah,
1997, 1999, 2003; Zulkii et al., 2008; Sia and Abdullah, 2011,
2012; Alias et al., 2012; Hakimi and Abdullah, 2013; Mustapha
and Abdullah, 2013).
The study area is located onshore in the northwest part of Sarawak which is situated in the low-lying coastal plain between the
Mukah and the Balingian Rivers of Sarawak, Malaysia (Fig. 1b).
Offshore area is currently active petroleum exploration area and
thus an evaluation of source rock quality in the onshore in the

32
M.H. Hakimi et al. / Marine and Petroleum Geology 48 (2013) 31e46
Figure 1. (a) Location map showing the coal bearing Tertiary basins in Malaysia. (b) Geological map of Balingian and Mukah coalelds and surrounding areas, Sarawak. Study area shown by box. (c) Map showing distribution of the
Balingian and Liang Formations around Balingian and Mukah coalelds, northwest Sarawak, Malaysia, and location of sampling points.

M.H. Hakimi et al. / Marine and Petroleum Geology 48 (2013) 31e46

Sarawak will contribute in such exploration activity. The current


study is focused on the Late Miocene to Late Pliocene coals in the
Mukah and Balingian coalelds (Fig. 1b and c). Despite the importance of these coalelds, little is known about the coal. Sia and
Abdullah (2011) worked on the concentration and association of
minor and trace elements in Mukah coal, with emphasis on the
potentially hazardous trace elements. Sia and Abdullah (2012)
published on geochemical and petrographic characteristics of
low-rank Balingian coals and their implications on depositional
conditions and thermal maturity.
The characterization of the coal using the integration of conventional geochemical and coal petrological approaches is the main
objective of this study with the purpose to investigate in detail the
hydrocarbon generation potential of the Tertiary coals from
northwest Sarawak and their depositional conditions. The organic
petrological and organic geochemical data involved total organic
carbon (TOC) content, pyrolysis yields, proximate data, bitumen
extraction, maceral composition and vitrinite reectance
measurements.
2. Geologic setting
The Mukah and Balingian coalelds are located in the low-lying
coastal plain between the Mukah and the Balingian Rivers of Sarawak, covering an area of approximately 710 km2 in Sarawak
(Fig. 1bec). The Mukah coaleld is underlain by the Balingian

33

Formation of Early Miocene age, which is in turn unconformably


overlain by the Begrih Formation of Early Pliocene age (Fig. 2a). The
contact between these two formations is marked by a wedge of
basal conglomerate, called Begrih Conglomerate (Wolfenden,
1960). The Balingian Formation is composed of mudstones,
shales, siltstones, sandstones, tufte, and coals. Foraminifera
identied by Visser and Crew (unpublished), such as Ammodiscus
sp., Glomospira sp., Haplophragmoides sp., and Trochammina sp.,
suggest a brackish-water depositional environment. The coaleld
contains 13 coal seams, comprising 5 major (well developed, with
economic potential) and 8 minor seams (Fig. 2a). The coal seams are
usually between 1 and 2 m thick, with a cumulative coal thickness
up to 16 m. The coal is characterized by a high amount of huminite
(89e99%), low to moderate amounts of liptinite (<1e9%), and trace
to low amounts of inertinite (<5%), on a mineral matter-free basis.
The coal is of lignite rank, with high total moisture, and low total
sulphur content, and ash yield (Sia and Dorani, unpublished). In
contrast, the accumulation of the Balingian coal took place in the
Liang Formation during Upper Pliocene. The Liang Formation is
unconformably underlain by the Lower Pliocene Begrih Formation
in the north and by the Eocene Belaga Formation in the south
(Figs. 1c and 2b) (Hutchison, 2005). The Liang Formation has a
thickness of approximately 950 m (Fig. 2b), and is made up of thick
and massive clays, sands, tufte, coal seams and gravel lenses. The
fauna identied in the coal zone include Ammodiscus sp., Glomospira sp., Haplophragmoides sp., and Trochammina sp., which

Figure 2. Generalized stratigraphic column of the Tertiary coal-bearing Balingian and Formations in the Mukah (a) and Balingian (b) coalelds.

34

M.H. Hakimi et al. / Marine and Petroleum Geology 48 (2013) 31e46

suggest a brackish-water environment of deposition. Outside the


coal zone, however, the sediments were deposited in a very
shallow, near-shore type of marine deposition (Hutchison, 2005).
Four coal seams with an accumulated thickness of 20.42 m have
been identied in this study (Fig. 2b). The coal has been classied as
Subbituminous C (Sia and Abdullah, 2012) and represents the only
minable lignite deposit reported in the country.

The Begrih Formation consists of conglomerates, conglomeratic


sandstones, mudstones, shales, tufte and also a coal seam. The
formation contains mixed marine and brackish-water fauna south,
suggesting depositional conditions that were probably predominantly littoral (Hutchison, 2005). However, towards the north and
east, pure marine fossil such as Ammobaculites sp., Bolivia spp.,
Cibicides sp., Elphidium sp., Frondicularia sp., Rotalia sp., Textularia

Table 1
Results of pyrolysis analysis with calculated parameters and proximate analysis (as-received basis) of the Mukah and Balingian coals in the northwest Sarawak.
Coalelds

Balingian
coaleld

Sampling
sites

Depth
(m)

Sample
ID

Lithology

BO1

Mine
Mine
Mine
Mine
Mine
Mine
Mine
Mine
Mine
Mine
Mine
Mine
Mine
Mine
Mine
Mine

BO1-1
BO1-2
BO1-3
BO1-4
BO1-5
BO1-6
BO1-7
BO2-1
BO2-4
BO2-5
BO3-1
BO3-2
BO3-2A
BO3-3
BO3-4
BO3-5

Coal
Coal
Coal
Coal
Coal
Coal
Coal
Coal
Coal
Coal
Coal
Coal
Coal
Coal
Coal
Coaly
sediments
Coal
Coal
Coal
Coal
Coal
Coal
Coal
Coal
Coal
Coal
Coaly
sediments
Coal
Coal
Coal
Coal
Coal
Coaly
sediments
Coal
Coal
Coal
Coal
Coal
Coal
Coal
Coal
Coal
Coal
Coal
Coal

BO2

BO3

ML46

Mukah
coaleld

MO1

MO2

MO3
MO4

MO5

O37
O38
O39
O43
O45
O46

face
face
face
face
face
face
face
face
face
face
face
face
face
face
face
face

Mine face
Mine face
12.11 m
14.11 m
13.11 m
11.11 m
10.11 m
38.3 m
37.30 m
Mine face
Mine face

BO3-6
BO3-9
ML43A-3
ML46A-1
ML46A-2
ML46A-4
ML46A-5
ML46B-1
ML46B-2
MO1-1
MO1-P

Mine
Mine
Mine
Mine
Mine
Mine

face
face
face
face
face
face

MO2-1
MO2-2
MO2-3
MO3-2
MO3-3
MO4-2

Mine face
Mine face
Mine face
Mine face
Outcrop
Outcrop
Outcrop
Outcrop
Outcrop
Outcrop
Outcrop
Outcrop

MO4-3
MO5-1
MO5-2
MO5-3
O37-1
O37-2
O38-1
O39-1
O43-B
O43C
O45
O46-B

S1: Volatile hydrocarbon (HC) content, mg HC/g rock.


S2: Remaining HC generative potential, mg HC/g rock.
S3: Carbon dioxide yield, mg CO2/g rock.
HI: Hydrogen Index S2  100/TOC, mg HC/g.
OI: Oxygen Index S3  100/TOC, mg CO2/g TOC.
Tmax: Temperature at maximum of S2 peak.
PI: Production Index S1/(S1 S2).
TOC: Organic Carbon, wt %.
MI: Total moisture content, wt % (as-received basis).
VM: Volatile Matter content, wt % (as-received basis).
FC: Fixed Carbon, wt % (as-received basis).
AS: Ash content, wt % (as-received basis).

Pyrolysis data
S1
(mg/g)

S2
(mg/g)

S3
(mg/g)

Tmax
( C)

S2/S3

HI

OI

PI

TOC
wt%

Proximate analysis
MI
wt%

VM
wt%

FC
wt%

AS
wt%

7.8
5.5
7.5
8.3
10.8
5.9
5.9
4.8
7.8
14.6
9.5
10.5
9.2
9.0
14.8
2.0

75.9
99.9
68.0
71.3
83.0
53.0
60.5
54.2
60.7
79.2
79.1
74.8
73.7
75.7
85.1
24.4

9.4
7.2
12.0
8.7
8.3
11.8
11.5
11.9
12.0
12.0
9.9
9.7
10.3
9.8
14.1
4.6

417
417
413
412
409
413
413
411
402
381
406
407
403
405
399
423

8.1
13.9
5.7
8.2
10.0
4.5
5.3
4.6
5.1
6.6
7.9
7.7
7.2
7.7
6.0
5.3

183
274
159
163
186
151
170
153
169
177
182
175
175
163
190
195

23
20
28
20
19
34
32
33
34
27
23
23
24
21
31
37

0.09
0.05
0.10
0.10
0.11
0.10
0.09
0.08
0.11
0.16
0.11
0.12
0.11
0.11
0.15
0.08

41.4
36.4
42.8
43.7
44.6
35.0
35.5
35.5
35.9
44.7
43.4
42.7
42.1
46.4
44.7
12.5

25.0
17.4
15.3
25.2
23.8
25.4
e
19.8
22.8
18.9
25.1
16.4
e
15.4
18.5
19.9

36.3
41.1
49.3
38.0
36.5
35.4
e
38.7
37.6
40.8
38.7
44.2
e
41.5
39.8
42.3

33.5
40.1
32.3
34.2
35.8
34.4
e
40.4
36.6
37.2
34.7
37.2
e
40.1
37.4
35.7

5.2
1.4
3.1
2.6
3.9
4.8
e
1.1
3.0
3.1
1.5
2.2
e
3.0
4.3
e

13.5
19.7
7.2
13.0
7.8
10.8
8.9
5.7
4.7
3.6
2.9

80.4
95.7
76.3
104.7
77.3
79.0
71.6
71.4
63.7
58.5
45.6

14.4
13.9
9.4
12.0
12.3
9.0
13.8
9.4
9.2
8.7
4.2

402
398
407
407
408
406
402
419
414
421
428

5.6
6.9
8.1
8.7
6.3
8.8
5.2
7.6
6.9
6.7
10.9

188
208
164
232
180
182
200
181
161
237
257

34
30
20
27
29
21
39
24
23
35
23

0.14
0.17
0.09
0.11
0.09
0.12
0.11
0.07
0.07
0.06
0.06

42.7
46.1
46.5
45
42.9
43.3
35.9
39.4
39.5
24.7
17.7

28.2
21.9
27.1
31.7
32.5
31.8
29.9
26.7
30.5
20.6
23.7

37.8
41.0
37.0
35.6
36.9
33.9
37.6
37.4
31.1
25.9
19.0

30.3
34.6
33.0
30.0
27.5
29.4
30.3
31.0
32.4
20.4
9.1

3.7
2.5
2.9
2.7
3.1
4.9
2.2
4.9
6.0
33.1
48.2

3.6
3.9
4.7
2.3
2.5
0.5

81.0
75.4
93.0
49.4
64.4
17.3

9.5
8.0
8.9
5.6
6.7
2.9

417
420
415
424
420
427

8.5
9.4
10.4
8.8
9.6
6.0

191
221
236
183
206
141

22
23
23
21
21
24

0.04
0.05
0.05
0.05
0.04
0.03

42.4
34.1
39.4
27.0
31.3
12.3

20.5
11.4
11.4
12.5
17.5
7.4

33.8
27.9
33.7
20.1
27.0
14.5

42.6
27.4
34.2
20.9
30.9
8.7

3.1
33.3
20.7
46.5
24.6
69.4

2.2
5.6
3.6
6.5
13.6
2.9
4.7
4.2
1.8
2.9
5.6
4.2

70.3
83.5
57.2
90.0
161.3
69.3
89.8
79.1
28.9
64.3
113.2
87.3

9.9
10.7
4.6
11.6
7.3
6.4
7.8
8.1
11.8
10.2
8.5
8.6

412
417
426
418
402
418
418
411
415
415
408
414

7.1
7.8
12.4
7.8
22.1
10.8
11.5
9.8
2.4
6.3
13.3
10.2

152
229
235
206
289
271
233
186
90
166
247
232

21
29
19
26
13
25
20
19
37
26
19
23

0.03
0.06
0.06
0.07
0.08
0.04
0.05
0.05
0.06
0.04
0.05
0.05

46.2
36.5
24.4
43.8
55.9
25.6
38.6
42.5
31.9
38.7
45.8
37.7

17.8
23.2
e
21.3
15.4
13.8
12.0
18.1
12.7
24.7
12.9
18.1

36.1
35.9
e
36.9
41.0
26.4
32.6
34.6
37.5
32.6
37.6
33.5

40.7
35.7
e
37.3
36.4
22.0
35.4
44.1
43.4
40.2
44.3
35.1

5.4
5.2
e
4.5
7.2
37.8
20.0
3.2
6.4
2.5
5.2
13.3

M.H. Hakimi et al. / Marine and Petroleum Geology 48 (2013) 31e46

35

Organic Petrology for huminite (Sykorova et al., 2005), liptinite


(Taylor et al., 1998) and inertinite (ICCP, 2001) nomenclature. The
reectance measurements were performed under a monochromatic light of 546 nm using a Leica DM6000M microscope and
an optical sapphire glass standard having a reectance of 0.589% in
oil immersion, following the procedures outlined by Taylor et al.
(1998). The rank was determined by measuring the random
reectance on huminite and the values reported were an average of
at least 100 measurements per sample.

sp., and Triloculina sp., replaces the mixed fauna, which indicates an
increasing marine inuence (Visser and Crew, unpublished). The
Belaga Formation is a highly deformed deep-water turbidity deposit of the Upper Cretaceous to the Middle Eocene age (Hutchison,
2005).
3. Sampling and methodology
A total of 45 samples were picked up from fteen sampling
sites (eleven sampling sites in the Mukah coaleld and four
sampling sites from Balingian coaleld; see Table 1 and Fig. 2),
applying the bench-by-bench channel sampling method. The
samples were collected from coal seams within the Balingian and
Liang Formations (Fig. 2). The sampling interval was decided on
the basis of changes in coal lithotypes, with each sample representing a single bench sample with a bench thickness of not more
than 1 m.

3.2. Organic geochemical analyses


Geochemical analyses include proximate and pyrolysis analyses,
as well as bitumen extraction, were performed at the Geochemistry
Laboratories of the Department of Geology in the University of
Malaya.
Part of the samples was ground to <150 mm and analysed on a
Diamond ThermogravimetriceDifferential Thermal Analyser (TGe
DTA). Proximate analysis carried out to determine moisture, ash,
volatile matter and xed carbon contents. Proximate analysis was
followed ASTM standard (1990).
All of the samples were screened by Source Rock Analyzer (SRA).
The collected samples were crushed into ne powder (<150 mm)
and analysed using (SRA-Weatherford)-TOC/TPH instrument
(equivalent to Rock-Eval equipment). Parameters measured are TOC
content and S1, S2, S3 pyrolysis yields and temperature of maximum

3.1. Petrographic analysis


For maceral analyses, the coal samples were crushed to a
maximum particle size of 1 mm, mounted in epoxy resin and polished. The maceral analyses were performed on a Leica DM6000M
microscope in monochromatic and UV light illumination on 1000
points. The maceral description used in this study follows the terminology developed by the International Committee for Coal and

Table 2
Random huminite reectance (R%), maceral composition (mineral free, %), and petrographic facies indices of the studied Mukah and Balingian coals.
Coalelds

Balingian
coaleld

Sampling
sites

Sample
ID

R%

BO1

BO1-1
BO1-2
BO1-3
BO1-5
BO1-6
BO2-1
BO2-5
BO2-4
BO3-1
BO3-2
BO3-3
BO3-4
BO3-6
BO3-9
ML46A-4
ML46A-5
ML46A-7
ML46B-1
ML46B-2
MO1-1
MO2-3
MO3-2
MO3-3
MO5-1
MO5-3
O37-1
O37-2
O38-1
O39-1
O43C
O46B

0.30
0.36
0.29
0.32
0.26
0.35
0.35
0.32
0.30
0.30
0.27
0.35
0.28
0.32
0.32
0.34
0.34
0.32
0.33
0.38
0.37
0.38
0.40
0.36
0.37
0.38
0.39
0.39
0.38
0.36
0.38
0.34

BO2

BO3

ML46

Mukah
coaleld

MO2
MO2
MO3
MO5
O37
O38
O39
O43
O46

Mean

Huminite (%)

Liptinite (%)

Inertinite (%)

Tx

Dh

Ch

Total

Sp

Cu

Rs

Ld

Sub

Ex

Total

Fg

Idt

Sf

Ma

Total

0
47
0
0
0
0
0
0
1
0
1
0
0
1
0
0
0
0
0
0
4
19
0
0
0
0
1
0
0
0
0
3

47
19
93
24
44
2
5
1
27
8
74
9
18
60
5
10
11
4
10
14
33
27
55
7
41
18
8
14
15
15
2
23

38
1
2
53
46
87
74
83
40
66
17
52
74
31
77
74
64
68
52
76
44
42
25
82
52
29
79
69
71
67
75
55

1
28
0
1
0
5
1
1
21
1
2
4
0
1
6
2
1
1
4
1
10
1
3
0
0
18
6
1
3
3
0
4

86
95
95
78
90
94
80
85
89
75
94
65
92
93
88
86
76
73
66
91
91
89
83
89
93
65
94
84
89
85
77
85

0
0
1
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
1
0
1

3
2
1
0
0
0
2
1
1
0
0
0
0
1
2
3
1
1
0
0
2
2
1
1
0
0
0
2
2
3
12
2

3
0
1
0
1
2
3
1
0
2
1
0
1
0
1
3
1
19
0
6
1
2
1
2
0
1
2
3
0
3
8
2

2
0
0
2
1
1
1
1
5
0
1
1
1
0
1
1
1
1
1
0
1
1
1
0
0
1
2
1
1
1
1
1

3
1
1
14
1
1
12
9
2
22
3
32
5
4
6
4
18
3
30
1
2
0
6
0
3
31
1
3
6
3
0
7

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
3
4
16
3
4
18
12
8
24
5
33
7
5
10
11
21
24
31
7
6
5
9
3
3
33
5
9
9
11
21
12

0
0
0
1
0
1
1
0
0
0
0
0
0
1
0
1
1
1
1
0
1
0
1
3
2
0
0
1
1
1
0
1

1
1
1
1
3
0
0
1
1
0
0
0
0
0
1
2
1
1
0
0
0
3
0
2
1
2
1
2
1
1
1
1

1
0
0
2
2
0
0
0
1
1
1
1
0
0
0
0
0
1
1
1
2
2
4
1
1
0
0
1
0
0
1
1

1
0
0
3
2
1
1
2
1
0
0
1
1
1
1
0
1
0
1
1
0
1
3
2
0
0
0
3
0
2
0
1

0
1
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

3
2
1
6
7
2
2
3
3
1
1
2
1
2
2
3
3
3
3
2
3
6
8
8
4
2
1
7
2
4
2
4

R%: Huminite reectance.


Huminite e Tx: Textinite; U: Ulminite; Dh: Detrohuminite; Ch: Corpohuminite.
Liptinite e Sp: Sporinite; Cu: Cutinite; Rs: Resinite; Ld: Liptodetrinite; Sub: Suberinite; Ex: Exsudatinite.
Inertinite e Fg: Funginite; Idt: Inertodetrinite; F: Fusinite; Sf: Semifusinite; Ma: Macrinite; TPI: Tissue Preservation Index; GI: Gelication Index.
TPI (telohuminite semifusinite)/(detrohuminite macrinite inertodetrinite).
GI huminite/inertinite.

TPI

GI

1.23
33.50
31.00
0.50
0.94
0.04
0.08
0.04
0.71
0.12
4.41
0.19
0.26
2.00
0.08
0.13
0.18
0.06
0.21
0.20
0.84
1.04
2.32
0.11
0.77
0.58
0.11
0.24
0.21
0.25
0.03
2.66

28.7
96.0
95.0
13.0
12.9
47.0
40.0
28.3
29.7
75.0
94.0
32.5
92.0
46.5
44.0
28.7
25.3
24.3
22.0
45.5
30.3
14.8
10.4
11.1
23.3
32.5
94.0
12.0
44.5
21.3
38.5
40.4

36

M.H. Hakimi et al. / Marine and Petroleum Geology 48 (2013) 31e46

S2 pyrolysis yield (Tmax). Hydrogen (HI), oxygen (OI) and production


(PI) Indies were calculated (Table 1).
The soluble organic matter (bitumen) was extracted from pulverized coal samples (<150 mm) using a Soxhlet apparatus for 72 h
using an azeotropic mixture of dichloromethane (DCM) and
methanol (CH3OH) (93:7 v/v). The fractions of the soluble organic
matter were separated into saturated hydrocarbons, aromatic hydrocarbons and NSO compounds by liquid column chromatography
over silica gel and aluminium oxide.
4. Results and discussion
4.1. Petrographic characteristics of coal
4.1.1. Huminite reectance measurements
The huminite reectance measurements of ulminite are given in
Table 2. The mean random reectance measurement of ulminite
maceral of the studied coals varied from 0.26 to 0.39%, indicating
the immature nature of the materials.

4.1.2. Maceral composition


The maceral content of the analysed coals is reported in Table 2
(vol. %, mineral matter free) and illustrated in Figures 3 and 4. The
coal samples are classied as humic coal (Fig. 5) and dominated by
huminite (65e95%), with low to high amounts of liptinite (3e33%)
and low amounts of inertinite (1e8%).
Huminite group is the most abundant macerals in the studied
coals and range from 65 to 96%, in total three sets of samples and
were occurred in approximately equal proportions of telohuminite
(ulminite/textinite) and detrohuminite (Table 2). Generally, individual samples are dominated by telohuminite and detrohuminite
as subgroup of Huminite group or the other, indicative of the
relative degree of degradation and decomposition of the original
peat material (Hackley et al., 2007). The detrohuminite (Fig. 3e)
content varied greatly in the studied coal, ranging from 1 to 87%
with an average of 55% (Table 2). The ulminite content is also varied
greatly in the studied coals (Fig. 3a, b and d), ranging from 1 to 93%
(Table 2). Textinite (Fig. 3c and f) was present in low concentrations
in most of the studied samples (1%), with the exceptions of only

Figure 3. Photomicrograph of (a) ulminite (U) associated with resinite (Rs) and funginite (Fg), inertodetrinite (Idt) and semifusinite (Sf); (b) ulminite (U) associated with sporinite
(Sp) and funginite (Fg); (c) the presence of textinite (Tx), ulminite (U), phlobaphinite (Ph) of the corpohuminite maceral and porigelinite (Pg); (d) the presence of ulminite (U),
phlobaphinite (Ph) of the corpohuminite maceral associated with resinite; (e) funginite (Fg) in the matrix of attrinite of the detrohuminite (De) and ulminite (U); (f) exsudatinite
(Ex) in the textinite (Tx). All photos were taken under reected light examination.

M.H. Hakimi et al. / Marine and Petroleum Geology 48 (2013) 31e46

37

Figure 4. Photomicrograph of (a) yellow to orange uorescing suberinite (Sb) associated with sporinite; (b) Bright yellow uorescing resinite (Rs) associated with cutinite (Cu) and
liptodetrinite (Ld); (c) yellow to orange uorescing cutinite (Cu); (d) yellow uorescing cutinite (Cu) associated with sporinite (Sp) and greenish uorescing uorinite (Fi); (e) yellow
to greenish uorescing uorinite (Fi) associated with sporinite (Sp) and cutinite (Cu); (f) yellow to greenish uorescing uorinite mega-sporinite (Sp) associated with liptodetrinite
(Ld). All photos were taken under reected UV light examination. (For interpretation of the references to colour in this gure legend, the reader is referred to the web version of this
article.)

samples MO2-3, BO1-2 and MO3-2 having a relatively high content


of textinite, at 4%, 19% and 48%, respectively (Table 2). The coals
contained low amounts of corpohuminite, usually below 6%
(Table 2), except for MO2-3, O37-1, BO3-1 and BO1-2 samples,
which contained 10%, 18%, 21% and 28% of corpohuminite respectively (Table 2). Corpohuminite occurred mainly as cell llings in
textinite (Fig. 3c and d).
Liptinite group is containing signicant amounts of terrestrial
liptinite macerals and ranging from 3 to 33% (Table 2). The liptinite
macerals were identied by their different nature of uorescence in
ultraviolet light (Fig. 4). The common liptinite macerals in the coals
are sporinite, cutinite, resinite, suberinite and liptodetrinite, while
the unstructured liptinites included uorinite and exsudatinite.
Higher liptinite content (>10%) was recorded in samples BO1-1,
BO1-5, BO2-5, BO2-4, BO3-2, BO3-4, ML46A-5, ML46A-7, ML46B1, ML46B-2, O37-1, O38-1, 046B, and O43C, with liptinite content
as high as 31e33% (Table 2). High liptinite content in these samples

was due mainly to the presence of high amounts of suberinite (up


to 33%), and also resinite (Table 2). Suberinite content varied greatly
from non-existent to 32% (Table 2), which it occurred as cell wall
tissues and was characterized by a yellow to brownish-yellow
uorescence (Fig. 4a). Resinite appeared mostly as an isolated
globular bodies (Fig. 4b), but there were also some cell-lling in
textinite (Sia and Abdullah, 2012). The maceral exists as inllings of
textinite cells and points towards the peat contribution from
coniferous vegetation (Stefanova et al., 2013). The contents of
cutinite and liptodetrinite were low, ranging from 0 to 12% and 0 to
5%, respectively (Fig. 4bed and Table 2). Liptodetrinite in the
studied coal was often present was as nely dispersed in a
groundmass of attrinite (Fig. 4b). Other liptinite macerals such as
sporinite, exsudatinite and uorinite had also been observed in the
studied coal, but they were either absent or were present in trace
amounts (Table 2). Exsudatinite in the studied coal appeared as
crack llings and was yellow in uorescent light (Fig. 3f). Fluorinite

38

M.H. Hakimi et al. / Marine and Petroleum Geology 48 (2013) 31e46

Figure 5. Ternary diagram of maceral group composition (huminiteeliptiniteeinertinite) for analysed Mukah and Balingian coals.

in the studied samples was always associated with cutinite and


characterized by a greenish-yellow in uorescent light (Fig. 4d and
e). Sporinite in the studied coal was appeared as micro and megasporinite and associated with cutinite (Fig. 4def).
Inertinite group was also widely detected in the studied coals
with low amount ranging from 1 to 8% (Table 2). The inertinite
macerals identied were semifusinite, fusinite, inertodetrinite,
macrinite and funginite. The inertodetrinite, semifusinite and
funginite are representing major macerals of the inertinite (1e3%),
whereas fusinite and macrinite are representing minor macerals
(<2% and <1%, respectively) as shown in Table 2. Fungus, seen as
the maceral funginite (Fig. 3a, b and e), is known to play a role in the
development of degraded maceral forms (Hower et al., 2010, 2011;
OKeefe and Hower, 2011).
4.2. Geochemical characteristics of coal
4.2.1. Molecular composition
The amounts of Extractable Organic Matter (EOM) together with
the relative proportions of saturated and aromatic hydrocarbons
and NSO (nitrogen/sulphur/oxygen) compounds have been

calculated and tabulated (Table 3). The EOM yields from the coals of
the Mukah and Balingian coalelds range from 25724.9 to
92143.7 ppm. All the extracts are mainly composed of heterocompounds (NSO) in the range of 65.7e95.1% (Table 3), which is
common for coals. In most of the extracts of analysed coals, aromatic hydrocarbon (aromatics; 2.9%e28.1%) are more abundant
than saturated hydrocarbons (alkanes; 2.0%e10.0%). The variation
in the alkane content has been related to the origin of plant material and to the intensity of biochemical degradation of the plant


material (Zivoti
c et al., 2008). The saturated and aromatic fractions
together create the petroleum-like hydrocarbon fraction; thus, the
sum of these two fractions is referred as hydrocarbons (HCs). Since
the hydrocarbon portion of the bitumen extracted from sediment is
the petroleum-like portion, it is used as an important parameter in
the source-rock evaluation (Philippi, 1957; Baker, 1972). In this
respect, the Mukah and Balingian coals are likely to be prolic
petroleum sources where abundant gas with limited oils may be
expected to generate. This is suggested by relatively high hydrocarbon fractions (4.9%e34.3%; see Table 3) and moderately saturated hydrocarbon proportions (2.0%e10.0%). The hydrocarbon
generative potential of a source rock can also be estimated from
plots of TOC content versus extractable organic matter (EOM) and
hydrocarbon yields (Fig. 6). The plots show that the studied coal
samples are viable source rocks for signicant gas and limited oilgeneration potential (Fig. 6) as supported by the TOC content and
pyrolysis S2/S3 yields (Fig. 7).
4.2.2. Organic geochemical analyses
Total organic carbon (TOC) content and pyrolysis data were
performed to character the organic content, hydrocarbon potential
of the organic matter and its thermal maturity level. Total organic
carbon (TOC) analysis demonstrated high TOC values of the coals
and coaly sediments from Mukah and Balingian coaleld (Table 1).
Samples from Mukah and Balingian coaleld are characterized by
very similar TOC values with slightly higher TOC values in the
Mukah samples (Table 1). The content of hydrocarbon yield (S2)
generated during pyrolysis is a useful parameter to evaluate the
generation potential of source rocks (Peters, 1986; Bordenave,
1993). The Mukah and Balingian samples were characterised by
pyrolysis S2 yield values in the range of 17.3e161.3 mg HC/g rock
(Table 1). These TOC contents and pyrolysis S2 yield values meet the
accepted standards of a source with very good to excellent source
rock potential (Fig. 8) as classied by Peters and Cassa (1994).
Hydrogen index (HI) and oxygen index (OI) of the studied samples
were calculated and determined in the range of 90e289 mg HC/g

Table 3
Bulk geochemical results of soluble organic matter (EOM) yields (ppm), relative proportions of saturated and aromatic hydrocarbons and NSO compounds of the EOM (in wt %)
of selected Mukah and Balingian coals.
Coalelds

Samples
ID

Bitumen extraction and chromatographic


fractions (ppm of whole rocks)
EOM

Sat.

Aro.

NSO

HC

Sat./EOM

Aro./EOM

NSO/EOM

HC

Balingian
coaleld

BO1-5
BO1-6
BO3-9
ML46B-2
MO1-1
MO3-3
O38-1
O46B

47405.3
45202.5
92143.7
31857.7
25724.9
29333.3
47260.8
47663.7

4724.00
1459.32
1813.14
1954.5
1590.76
2120.48
2297.00
2008.30

3283.18
4377.96
2629.05
5863.4
7226.62
2456.22
4881.13
6761.28

39398.2
39365.2
87701.5
24039.8
16907.6
24756.6
40082.7
38894.1

8007.18
5837.29
4442.19
7817.90
8817.38
4576.70
7178.13
8769.58

10.0
3.2
2.0
6.1
6.2
7.2
4.9
4.2

10.9
9.7
2.9
18.4
28.1
8.4
10.3
14.2

79.1
87.1
95.1
75.5
65.7
84.4
84.8
81.6

20.9
12.9
4.9
24.5
34.3
15.6
15.2
18.4

Mukah
coaleld

EOM Extractable organic matter (Bitumen extraction).


Sat Saturated hydrocarbons.
Aro Aromatic hydrocarbons.
NSO Nitrogen, Sulphur, Oxygen components.
HC Hydrocarbon fractions (Saturated Aromatic).

Chromatographic fractions of
Bitumen extraction (EOM wt%)

TOC
(wt.%)

HC/TOC
(mg/g TOC)

44.6
35.0
46.1
39.5
24.7
31.3
38.6
37.7

179.5
166.8
96.4
197.9
356.9
146.2
185.9
232.6

M.H. Hakimi et al. / Marine and Petroleum Geology 48 (2013) 31e46

39

Figure 6. Plots of TOC content versus bitumen extractions and hydrocarbon yields, showing source potential rating and hydrocarbon source-rock richness for the selected coals.

TOC and 13e39 mg CO2/g TOC, respectively (Table 1). In addition,


Tmax value which represents the temperature at the point where S2
peak is the maximum is also determined (Espitali et al., 1977),
whereby Tmax values ranging from 381 to 428  C (Table 1), indicating the immature organic matter as agreement with huminite
reectance values (Fig. 9) and supported by hydrocarbon extraction
yields (Fig. 10). A diagram was constructed based of pyrolysis data,
kerogen classication and thermal maturity using the HI versus
Tmax data as carried out by above cited authors. In general, the results show that the Balingian coals are dominated by Type III
kerogen while Mukah coals are dominated by Type II/III kerogens
(Fig. 11). In this respect, all the analysed samples are generally
plotted in the immature organic matter of Type III kerogen and
mixed Type II/III kerogens (Fig. 11). In comparison to the other
coalelds, Mukah coals are more capable to generate oil comparing

to Balingian coals. This is most probably due to the higher pyrolysis


S2/S3 yields of the Mukah coals (Fig. 7).
4.2.3. Proximate analysis and coal rank
Proximate geochemical analysis is carried out to determine the
rank and suitability of coals for various industries uses (ASTM,
1989) and to appreciate thermal maturity of the coal bearing sequences and individual coal seams (Ward, 2002). Results from
proximate analysis indicated that the coals from Mukah and Balingian coalelds generally are characterized by similar volatile
matter and xed carbon contents with higher values in the Balingian coal samples (Table 1). The volatile matter and xed carbon
contents indicate that the coals could be consider as lignite or
transitional to sub-bituminous C coals according to ASTM classication (Stach et al., 1982). This interval is also conrmed by

40

M.H. Hakimi et al. / Marine and Petroleum Geology 48 (2013) 31e46

inorganic matter, rather than to coalication degree (e.g., Stefanova


et al., 2013).
The total moisture contents obtained from the coal samples in
Balingian and Mukah coalelds are given in Table 1. In comparison
to other coalelds, coal samples from the Mukah coaleld are
characterized low total moisture contents, in the range of 7.4e
24.7 wt. %, whereas the Balingian coals are characterized by high
total moisture contents, in the range of 15.3e32.5 wt. %. The lower
total moisture content in the Mukah coals is explained by maturity
of the Mukah coals more mature comparing to the Balingian coals
as indicated by huminite reectance and pyrolysis Tmax data
(Fig. 13). The main explanation could be the depth of burial and age
of the analysed coals. The Mukah coals are older than Balingian
coals as concluded from stratigraphic sections (see Section 2).
4.3. Hydrocarbon generation potential

Figure 7. A plot of total organic carbon (TOC) versus S2/S3 yields, showing potential
hydrocarbon generative and type of the studied coals.

huminite reectance values, which vary in the range of 0.26 and


0.39%.
The ash content of the analysed coal samples from the Balingian
and Mukah coalelds (Sarawak basin) ranges between 1.1 wt% to
69.4 wt% (Table 1). Generally, the ash contents were low to moderate for the Balingian coals, ranging from 1.1 to 6.0 wt%, whereas
the Mukah coals are characterized by high ash contents in the range
of 2.5e69.4 wt. % (Table 1). In the literature, the presence of high
ash content is explained by different phenomenon. One is the peat
depositional environment where conditions are a periodic ooding


of the paleomire during deposition (Zivoti
c et al., 2013). Moreover,
Siavalas et al. (2009) reported the high ash contents from the Tertiary and Quaternary coal deposits in the Megalopolis Basin and
related to the high inorganic input during peat accumulation. The
high ash content of Mukah coals is explained by the high peat
inorganic input during peat accumulation as indicated by Sia and
Abdullah (2011). In addition, TOC values show good negative correlation with the ash contents (Fig. 12; R2 0.75) indicating that the
changes in TOC contents are due to the increased contribution of

Figure 8. Pyrolysis S2 versus total organic carbon (TOC) plot showing generative
source rock potential for the Tertiary coals from northwest Sarawak.

The type of the organic matter and content of hydrocarbons that


might be generated were appreciated based on organic geochemical and optical data. Different types of kerogen will produce
different type of hydrocarbons. Generally, Type I and Type II
kerogen characterizing marine and lacustrine type are considered
as the best kerogen capable to generate liquid hydrocarbons. Type
III kerogens that are mostly characterized by terrigenous organic
matter are considered as gas-prone. Several studies indicate that
there is a direct correlation between pyrolysis data and petroleum
generation potential (Bordenave et al., 1993; Hunt, 1996). The
thresholds for the source rock quality are based on HI values
dened by Peters and Cassa (1994): samples that contain a Type III
vitrinitic kerogen would be expected to generate gas with hydrogen
index <200 mg HC/g TOC whereas samples with hydrogen index
values higher than 200 mg HC/g TOC can generate oil although
their main generation products are gas and condensate. Moreover,
samples characterized by HI higher than 300 mg HC/g TOC can
generate oil (Bordenave et al., 1993; Hunt, 1996). In this study, for
the coal samples are determined HI values in the range of 90e
289 mg HC/g TOC, indicating that the coals can generate oil
although their main generation products are gas. This is supported
by the characteristics of the extracted bitumen and hydrocarbon
yields.
Petrographic observations were also performed and the data
obtained were compared with the results from pyrolysis. Microscopic observations of the studied coal samples indicate that the
coals are characterized by low to high amounts of liptinite (Table 2).
The signicant representation of liptinite macerals are sporinite,
suberinite, resinite, liptodetrinite and cutinite. Suberinite, a maceral that is almost exclusively found in Tertiary coals and only in a
few Mesozoic coals (Teichmller and Teichmller, 1982) was recognised as the most common liptinite maceral in this study
(Table 2). This suberinite maceral shows a close association with the
macerals sporinite, cutinite, exsudatinite and liptodetrinite. It is
most apparent that the oil-prone nature of the analysed coals is
predominantly attributed to the common occurrence of suberinite
and its associated macerals such as liptodetrinite, resinite, cutinite,
bituminite, sporinite and exsudatinite. Suberinite and cutinite are
abundant waxy oils might be the expected products (Stasiuk et al.,
2006), and where resinite and other liptinite macerals (e.g., liptodetrinite and sporinite) are likely the most productive sources in
the coals thus where the former is abundant naphthenic oil to
condensate might be expected (Petersen et al., 2013).
The relatively higher value of HI and liptinite content are also
supported by the presence of n-alkene/n-alkane doublets predominant in the open system pyrolysis gas chromatography (PyGC) for a coal samples (e.g., Wan Hasiah, 1999; Petersen et al., 2001;
Alias et al., 2012; Hakimi and Abdullah, 2013). The open system

M.H. Hakimi et al. / Marine and Petroleum Geology 48 (2013) 31e46

41

Figure 9. A plot of Tmax values versus huminite reectance (%Ro) values, showing good agreement between Tmax and huminite reectance (%Ro) data and generally an immature
stage for the Mukah and Balingian coals.

pyrolysis gas chromatography (Py-GC) of the analysed coals


(Fig. 14) depicts mixed kerogen ngerprints of predominantly nalkane/n-alkene doublets and aromatic compounds. In the Mukah
coal samples where the liptinite macerals content is higher than
35% and relatively high hydrogen index values (289 mg HC/g TOC)
are determined, the Py-GC is dominated by n-alkane/n-alkene
doublets extended beyond C30 are indicative for aliphatic-rich, oilprone nature of these macerals (Fig. 14a) (Petersen et al., 2001). In
contrast, in the Balingian coal samples characterized by relatively
low liptinite macerals content and low hydrogen index values, the
Py-GC also is displayed by n-alkane/n-alkene doublets extended
beyond C30. They are indicative for aromatic-rich, and with significant aliphatic compounds, suggest oil although their main generation products are gas (Fig. 14b).

4.4. Palaeoenvironmental conditions of peat-forming


The maceral composition of coals reects the organic source
materials which contributed to the accumulation of peat and
further provides information about the conditions during deposition (Kalkreuth et al., 1991). Maceral analysis measures the relative
proportions and interrelationships of various maceral groups
(Table 2). The diagnostic macerals and petrographic facies indices
derived from this analysis have been used as an indicator for the
palaeoenvironment of the coal-forming peat.
The high amount of huminite macerals with a general predominance of detrohuminite (Table 2), indicates overall oxygendecient depositional conditions in the peat-forming mires and
deposition in waterlogged conditions (wet forest) (Flores, 2002;

Figure 10. A plot of extract yield (mg HC/g TOC) versus percent of hydrocarbon in the total extracts (HC%), showing source rock potential and maturity level (modied after Powell,
1978).

42

M.H. Hakimi et al. / Marine and Petroleum Geology 48 (2013) 31e46

Figure 11. Plot of Hydrogen Index (HI) versus pyrolysis Tmax, showing kerogen quality and thermal maturity stages of the analysed coal samples in the northwest Sarawak.

Petersen et al., 2009; Erik, 2011). Detrohuminite has also been


considered to be derived from herbaceous, cellulose-rich wood
(Teichmller, 1989) and from poor preserved big woody plants
due to prolonged humication in slowly subsiding paleomires
(Diessel, 1992; Petersen et al., 2009; Sarez-Ruiz et al., 2012). This

suggests that the resulting maceral composition is also inuenced


by the degradational conditions in the mires as indicated by the
percent of funginite maceral (Fig. 3a, b and e) (e.g., Hower et al.,
2010, 2011; OKeefe and Hower, 2011). In contrast, the low
amount of detrohuminite in the studied coals was always

Figure 12. A plot of total organic carbon (TOC) versus ash contents of the analysed coal samples in the northwest Sarawak.

M.H. Hakimi et al. / Marine and Petroleum Geology 48 (2013) 31e46

43

Figure 13. Scatter plot showing the relationship between moisture and vitrinite/huminite reectance and pyrolysis Tmax.

accompanied by increase of ulminite and/or textinite and corpohuminite, indicating either increasing forest type mires or a lower
degree of gelication under relatively dry conditions (Sia and
Abdullah, 2012).
The relatively low inertinite content in the Balingian and Mukah
coals assumes low levels of peat re and/or oxidation occurred in
these mires (e.g. Scott and Glasspool, 2007). The presence of high
amounts of liptinite group macerals like suberinite, resinite and the
presence of cutinite also suggests an accumulation in a forest type
swamp (Ratanasthien et al., 1999; Erik, 2011).
The palaeoenvironment of the coal-forming peat has also been
interpreted using petrographic facies. In petrographic facies, the
petrographic composition of coal seams and petrography-based
facies indicators (gelication index (GI) and tissue preservation
index (TPI)) have been used to track the evolution of peatforming depositional environments (Calder et al., 1991; Diessel,
1986, 1992; Kalkreuth et al., 1991; Siavalas et al., 2009; Jasper


et al., 2010; Koukouzas et al., 2010; Zivoti
c et al., 2013 and

many others). The GI-TPI diagram was rstly proposed by


(Diessel, 1986) for high-rank Australian Permian coals. For low
rank Miocene and Jurassic coals, these indices have been modied by Kalkreuth et al. (1991). The GI and TPI are used in the
present study as they were modied by Sia and Abdullah (2012)
for low-rank Tertiary coals (Table 2). The petrographic data from
the present study indicate signicant fragmentation of the
organic matter in the Mukah and Balingian coals (Tissue Preservation Index, TPI < 1) and high gelication (Gelication Index,
GI > 1) for most of the coal samples (Fig. 15). The fragmentation
of the organic matter is mainly due to the herbaceous peatforming plants, and hence the accumulation mostly of soft tissues and trees were rare (Koukouzas et al., 2010 Jasper et al.,
2010). This is also due to a hydrodynamic level that favoured
the mechanical destruction of tissues during short-term transportation (Koukouzas et al., 2010).
For the coals with low TPI values (<1; Table 2) could be
assumed a high large scale destruction of wood in forested

44

M.H. Hakimi et al. / Marine and Petroleum Geology 48 (2013) 31e46

Figure 14. Open system pyrolysis gas chromatograms for representative samples from the: (a) Balingian coaleld (BO3-9) and (b) Mukah (O37-1) coalelds of the northwest
Sarawak.

swamps (Amijaya and Littke, 2005; Diessel, 1992) Respectively,


the organic matter preservation was low to extremely low
(Siavalas et al., 2009). Moreover, the low TPI values (TPI < 1;
Table 2) could imply signicant contribution gymnosperm vegetation is more resistant to decay because of high content of
resinous compounds. They protect tissue from microbial attack
(Mandic et al., 2008). In contrast, high TPI value (TPI > 1) was
determined for samples BO1-1, BO1-2, BO1-3, BO1-2, BO3-3 and
BO3-9, suggesting well-preserved plant tissues (textinite and


ulminite) (Zivoti
c et al., 2013). On the other hand, it was suggested
that tissue preservation depends mostly on the water level and
the climatic conditions during peat accumulation, rather than on

c et al., 2013).
the botanical properties of the vegetation (Zivoti
According to this study, higher GI values (>10), it could be
assumed that during peatication water column level was mod

erate to high (e.g., Zivoti
c et al., 2013) and represent predominantly topogenous mire conditions (Jasper et al., 2010). Moreover,
the high values of GI in the studied coals further suggest

gelication of plant tissues in continuous wet forest swamp


(Diessel, 1992; Sia and Abdullah, 2012) and could imply pro

nounced microbial activity (Zivoti
c et al., 2013).
TPI vs. GI diagram is shown in Figure 15 shows data of all the
samples. From it could be surmise that palaeomire was created by
herbaceous plants able to thrive in a marsh-wet forest swamp
environment. Nevertheless, marsh and forested swamp are
considered as a kind of minerotrophic mires (Amijaya and Littke,
2005). Coals originating from both of these sources usually
generate relatively high ash yield (Amijaya and Littke, 2005;
Diessel, 1992), which is consistent with the case for the studied
coals with ash contents in the range of 1.1e69.4 wt.%. In the present
case, the low TPI (TPI < 1) which is accompanied by relatively high
ash contents of the studied coal, could also be related to sedimentary environment with clastic contribution such as a lower
deltaic plain (Escobar and Martnez; 1997). This shows that the
interpretation as suggested by the Diessels diagram is valid for the
studied coals (Fig. 15).

M.H. Hakimi et al. / Marine and Petroleum Geology 48 (2013) 31e46

45

Figure 15. Diagram of TPI versus GI showing the paleodepositional environment of the Balingian and Mukah coals facies.

5. Conclusions
Organic petrographic and geochemical analyses were performed on the coal seams within the Balingian and Liang Formations in Mukah and Balingian coalelds of northwestern Sarawak.
Coal rank and petroleum generation potential as well as palaeoenvironment conditions using critical petrographic facies and
maceral compositions were studied. The data gave ground to
formulate the following conclusions:
(1) The organic matter is classied on pyrolysis HI versus Tmax
diagram the OM was determined as predominantly Type III
kerogen (gas-prone) grading into mixed Type IIeIII kerogens
(oil and gas-prone) as indicated by hydrogen index values (90e
289 mg HC/g TOC). This assumption is also supported by
macerals compositions, dominated by huminite, with low to
moderate amounts of liptinite.
(2) The geochemical classication of thermal maturity (coal rank)
based on proximate and huminite reectance values suggest
that the Mukah and Balingian coals are generally thermally
immature for hydrocarbon generation potential and range from
lignite to sub-bituminous C rank. In addition, the pyrolysis Tmax
data and hydrocarbon extraction yields conrm this attained
thermal maturity level.
(3) The studied coals are dominated by huminite (detrohuminite
and ulminite) with low amounts of inertinite, suggesting predominantly herbaceous plants in the paleomires preserved
under anaerobic deposition conditions with limited thermal
and oxidative tissues destruction.
(4) Most of the studied coal samples are characterized by low TPI
and high GI, suggesting a lower deltaic plain wet peat-swamp
depositional setting. They are also plotted on the marsh eld
of the Diessels diagram, which usually are characterised by

relatively high ash contents (Amijaya and Littke, 2005; Diessel,


1992), consistent with the observation for the studied coals.
Acknowledgements
The authors would like to sincerely thank Mr. Dorani J. and Mr.
Bakar J. of Sarawak Coal Resources Sdn. Bhd., who provided the
opportunity to access the sampling sites. The authors are most
grateful to the Department of Geology, University Malaya for
providing facilities to complete this research. The authors also
would like to express their gratitude to Mr. Wong, V.C. of the
Minerals and Geoscience Department Malaysia (Sabah) for eld
assistance and Ms. Jacinta John for arranging eld transportation.
We would like to sincerely thank an associate Editor Massimo
Zecchin and an anonymous reviewer for their careful and useful
comments that improved the revised manuscript. The study
received nancial support from the University of Malaya research
grants (Nos. PS438-2010A, RP002C-13AFR and RG145-11AFR).
References
Alias, F.L., Abdullah, W.H., Hakimi, M.H., Azhar, M.H., Kugler, R.L., 2012. Organic
geochemical characteristics and depositional environment of the Tertiary Tanjong Formation coals in the Pinangah area, onshore Sabah, Malaysia. International Journal of Coal Geology 104, 9e21.
Amijaya, H., Littke, R., 2005. Microfacies and depositional environment of Tertiary
Tanjung Enim low rank coal, South Sumatra Basin, Indonesia. International
Journal of Coal Geology 61, 197e221.
ASTM (American Society for Testing and Materials) D3176, 1989. Standard Practice
for Ultimate Analysis of Coal and Coke. In: Annual Book of ASTM Standards, vol.
05.05.
ASTM (American Society for Testing and Materials) D388, 1990. Standard Classication of Coals by Rank. In: Annual Book of ASTM Standards, vol. 05.05.
Baker, D.R., 1972. Organic geochemistry and Geological interpretations. Journal
Geo1ogical Education 20, 221e234.
Bordenave, M.L., 1993. Applied Petroleum Geochemistry. Editions Technip, Paris.

46

M.H. Hakimi et al. / Marine and Petroleum Geology 48 (2013) 31e46

Bordenave, M.L., Espitali, J., Leplat, P., Oudin, J.L., Vandenbroucke, M., 1993.
Screening techniques for source rock evaluation. In: Bordenave, M.L. (Ed.),
Applied Petroleum Geochemistry. Editions Technip, Paris, pp. 217e278.
Calder, J.H., Gibling, M.R., Mukopadhyay, P.K., 1991. Peat formation in a West phalian
B piedmont setting, Cumberland Basin, Nova Scotia: implications for the maceral based interpretation of rheotrophic and raised paleo-mires. Bulletin de la
Socit Gologique de France 162, 283e298.
Diessel, C.F.K., 1986. On the correlation between coal facies and depositional environments. In: Advances in the Study of the Sydney Basin, Proc. 20th Symposium. University of Newcastle, Australia, pp. 19e22.
Diessel, C.F.K., 1992. Coal-bearing Depositional Systems. Springer-Verlag, New York,
Berlin, p. 721, 3 540 52516 5.
EIA (Energy Information Administration), 2012. International Energy Outlook 2010.
Energy Information Adminstration, Washington DC, USA. http://204.14.135.140/
cfapps/ipdbproject/iedindex3.cfm?tid1&pid7&aid1&cidMY,&s
yid2010&eyid2010&unitTST.
Erik, N.Y., 2011. Hydrocarbon generation potential and MioceneePliocene paleoenvironments of the Kangal Basin (Central Anatolia, Turkey). Journal of Asian
Earth Sciences 42, 1146e1162.
Escobar, M., Martnez, M., 1997. Geoqumica Orgnica De La Formacin Los Cuervos
En San Pedro Del Ro, Estado Tchira, Venezuela. In: Proc. VIII Cong. Geol.
Venezolano, pp. 235e242.
Espitali, J., Laporte, J.L., Madec, M., Marquis, F., Leplat, P., Pauletand, J., et al., 1977.
Methode rapide de caracterisation des roches meres, de leur potential petrolier
et de leur degre devolution. Revue de lInstitut Francais du Petrole 32, 23e42.
Flores, D., 2002. Organic facies and depositional palaeoenvironment of lignites from
Rio Maior Basin (Portugal). International Journal of Coal Geology 48, 181e195.
Hackley, P.C., Warwick, P.D., Breland Jr., F.C., 2007. Organic petrology and coalbed
gas content, Wilcox Group (PaleoceneeEocene), northern Louisiana. International Journal of Coal Geology 71, 54e71.
Hakimi, M.H., Abdullah, W.H., 2013. Liquid hydrocarbon generation potential from
Tertiary Nyalau Formation coals in the onshore Sarawak, Eastern Malaysia. International Journal of Earth Sciences 102, 333e348.
Hower, J.C., OKeefe, J.M.K., Volk, T.J., Watt, M.A., 2010. Funginiteeresinite associations in coal. International Journal of Coal Geology 83, 64e72.
Hower, J.C., OKeefe, J.M.K., Eble, C.F., Raymond, A., Valentim, B., Volk, T.J.,
Richardson, A.R., Satterwhite, A.B., Hatch, R.S., Stucker, J.D., Watt, M.A., 2011.
Notes on the origin of inertinite macerals in coal: evidence for fungal and
arthropod transformations of degraded macerals. International Journal of Coal
Geology 86, 231e240.
Hunt, J.M., 1996. Petroleum Geochemistry and Geology, second ed. W.H. Freeman,
San Francisco.
Hutchison, C.S., 2005. Geology of North-west Borneo: Sarawak, Brunei and Sabah,
rst ed. Elsevier, New York, USA.
International Committee for Coal Petrology (ICCP), 2001. The new inertinite classication (ICCP System 1994). Fuel 80, 459e471.
Jasper, K., Hartkopf-Frder, C., Flajs, G., Littke, R., 2010. Evolution of Pennsylvanian
(Late Carboniferous) peat swamps of the Ruhr Basin, Germany: comparison of
palynological, coal petrographical and organic geochemical data. International
Journal of Coal Geology 83, 346e365.
Kalkreuth, W., Kotis, T., Papanicolaou, C., Kokkinakis, P., 1991. The geology and coal
petrology of a Miocene lignite prole at Meliadi Mine, Katerini, Greece. International Journal of Coal Geology 17, 51e67.
Koukouzas, N., Kalaitzidis, S.P., Ward, C.R., 2010. Organic petrographical, mineralogical and geochemical features of the Achlada and Mavropigi lignite deposits,
NW Macedonia, Greece. International Journal of Coal Geology 83, 387e395.
Mandic, O., Pavelic, D., Harzhauser, M., Zupanic, J., Reischenbacher, D.,
Sachenhofer, F.R., Tadej, N., Vranjkovic, A., 2008. Depositional history of the
Miocene Lake Sinj (Dinaride Lake System, Croatia): a long-lived hard-water lake
in a pull-apart tectonic setting. Journal of Paleolimnology 41, 431e452.
Mustapha, K.A., Abdullah, W.H., 2013. Petroleum source rock evaluation of the
Sebahat and Ganduman Formations, Dent Peninsula, Eastern Sabah, Malaysia.
Journal of Asian Earth Sciences. http://dx.doi.org/10.1016/j.jseaes.2012.12.003.
OKeefe, J.M.K., Hower, J.C., 2011. Revisiting Coos Bay, Oregon: a re-examination of
funginiteehuminite relationships in Eocene subbituminous coals. International
Journal of Coal Geology 85, 34e42.
Peters, K.E.,1986. Guidelines for evaluating petroleum source rock using programmed
pyrolysis. American Association of Petroleum Geologists Bulletin 70, 318e386.
Peters, K.E., Cassa, M.R., 1994. Applied source rock geochemistry. In: Magoon, L.B.,
Dow, W.G. (Eds.), The Petroleum System d From Source to Trap, American
Association of Petroleum Geologists, Memoir 60, pp. 93e120.
Petersen, H.I., Andersen, C., Anh, P.H., Bojesen-Koefoed, J.A., Nielsen, L.H., Nytoft, H.P.,
Rosenberg, P., Thanh, L., 2001. Petroleum potential of Oligocene lacustrine mudstones and coals at Dong Ho, Vietnam d an outcrop analogue to terrestrial source
rocks in the greater Song Hong Basin. Journal Asian Earth Sciences 19, 135e154.

Petersen, H.I., Lindstrm, S., Nytoft, H.P., Rosenberg, P., 2009. Composition, peatforming vegetation and kerogen parafnicity of Cenozoic coals: relationship
to variations in the petroleum generation potential (Hydrogen Index). International Journal of Coal Geology 78, 119e134.
Petersen, H.I., verland, J.A., Solbakk, T., Bojesen-Koefoed, J.A., Bjerager, M., 2013.
Unusual resinite-rich coals found in northeastern Greenland and along the
Norwegian coast: Petrographic and geochemical composition. International
Journal of Coal Geology 109e110, 58e76.
Philippi, G.T., 1957. Identication of oil source beds by chemical means. In: International Geo1ogical Congress, 20th, Mexico 1956, 3, pp. 25e38.
Ratanasthien, B., Kandharosa, W., Chompusri, S., Chartprasert, S., 1999. Liptinite in
coal and oil source rocks in northern Thailand. Journal of Asian Earth Sciences
17, 301e306.
Scott, A.C., Glasspool, I.J., 2007. Observations and experiments on the origin and
formation of inertinite group macerals. International Journal of Coal Geology 70
(1e3), 53e66.
Sia, G.S., Abdullah, W.H., 2011. Concentration and association of minor and trace
elements in Mukah coal from Sarawak, Malaysia, with emphasis on the
potentially hazardous trace elements. International Journal of Coal Geology 88,
179e193.
Sia, G.S., Abdullah, W.H., 2012. Geochemical and petrographical characteristics of
low-rank Balingian coal from Sarawak, Malaysia: its implications on depositional conditions and thermal maturity. International Journal of Coal Geology
96e97, 22e38.
Siavalas, G., Linou, M., Chatziapostolou, A., Kalaitzidis, S., Papaefthymiou, H.,
Christanis, K., 2009. Palaeoenvironment of Seam I in the Marathousa Lignite
Mine, Megalopolis Basin (Southern Greece). International Journal of Coal Geology 78, 233e248.
Stach, E., Mackowsky, M.T., Teichmller, M., Taylor, G.H., Chandra, D.,
Teichmller, R., 1982. Stachs Textbook of Coal Petrology. Borntraeger, Berlin.
Stasiuk, L.D., Goodarzi, F., Bagheri-Sadeghi, H., 2006. Petrology, rank and evidence
for petroleum generation, Upper Triassic to Middle Jurassic coals, Central Alborz
Region, Northern Iran. International Journal of Coal Geology 67, 249e258.
Stefanova, M., Kortenski, J., Zdravkov, A., Marinov, S., 2013. Paleoenvironmental
settings of the Soa lignite basin: insights from coal petrography and molecular
indicators. International Journal of Coal Geology 107, 45e61.
Sarez-Ruiz, I., Flores, D., Filho, J.G.M., Hackley, P.C., 2012. Review and update of the
applications of organic petrology: part 1, geological applications. International
Journal of Coal Geology 99, 54e112.
Sykorova, I., Pickel, W., Christanis, K., Wolf, M., Taylor, G.H., Flores, D., 2005. Classication of huminite e ICCP system 1994. International Journal of Coal Geology
62, 85e106.
Taylor, G.H., Teichmller, M., Davis, A., Diessel, C.F.K., Littke, R., Robert, P., 1998.
Organic Petrology. Gebruder Borntraeger, Berlin, Stuttgart, p. 704.
Teichmller, M., 1989. The genesis of coal from the viewpoint of coal petrology.
International Journal of Coal Geology 12, 1e87.
Teichmller, M., Teichmller, R., 1982. Fundamental of coal petrology. In: Stach, E.,
Mackowsky, M-Th., Teichmller, M., Taylor, G.H., Chandra, D., Teichmller, R.
(Eds.), Stachs Textbook of Coal Petrology, third ed. Gebrder Borntraeger,
Berlin, p. 535.
Wan Hasiah, A., 1997. Common liptinitic constituents of Tertiary coals from the
Bintulu and Merit-Pila Coaleld, Sarawak and their relation to oil generation
from coal. Geological Society of Malaysia Bulletin 41, 88e94.
Wan Hasiah, A., 1999. Oil-generating potential of Tertiary coals and other organicrich sediments of the Nyalau Formation, onshore Sarawak. Journal of Asian
Earth Sciences 17, 255e267.
Wan Hasiah, A., 2003. Coaly source rocks of NW Borneo: role of suberinite and
bituminite in oil generation and expulsion. Geological Society of Malaysia
Bulletin 47, 153e163.
Ward, C.R., 2002. Analysis and signicance of mineral matter in coal seams. International Journal of Coal Geology 50, 135e168.
Wolfenden, E.B., 1960. The geology and mineral resources of the Lower Rajang
Valley and adjoining areas, Sarawak. In: Memoir, 11. Geological Survey of British
Borneo, p. 167.
Zulkii, S., Awang, S.A.J., Kamal, R.M., Che, Aziz A., 2008. Hydrocarbon generation
potential of the coals and shales around the Eucalyptus Campsite area, Maliau
Basin, Sabah. Geological Society of Malaysia, Bulletin 54, 147e158.

Zivoti
c, D., Wehner, H., Cvetkovi
c, O., Jovan
ci
cevi
c, B., Gr
zeti
c, I., Scheeder, G.,


Vidal, A., Aleksandra Sajnovi
c, A., Ercegovac, M., Simi
c, V., 2008. Petrological,
organic geochemical and geochemical characteristics of coal from the Soko
mine, Serbia. International Journal of Coal Geology 73, 285e306.



Zivoti
c, D., Stojanovi
c, K., Gr
zeti
c, I., Jovan
ci
cevi
c, B., Cvetkovi
c, O., Sajnovi
c, A.,
Simi
c, V., Stojakovi
c, R., Scheeder, G., 2013. Petrological and geochemical
composition of lignite from the D eld, Kolubara basin (Serbia). International
Journal of Coal Geology 111, 5e22.

You might also like