You are on page 1of 26

A Geochemical Appraisal of Oil Generation in the

Taranaki Basin, New Zealand1


S. D. Killops,2 A. D. Woolhouse,3 R. J. Weston,3 and R. A. Cook2

ABSTRACT
From a basin-wide evaluation of organic geochemical data, it has been possible to characterize
and differentiate various source rock units and to
establish the genetic relationships of oils. Most of
the oils are primarily terrestrially sourced, and it is
possible to recognize varying contributions from
source rocks within the Paleogene Kapuni Group
and late Cretaceous Pakawau Group. This distinction is attributable to the rise to dominance of
angiosperms over gymnosperms in coastal plain
swamp communities by the Eocene. Apart from the
Maui family (i.e., Maui field, Maui-4, and Moki-1)
oils, the inferred relative contributions from the
main source rock types generally correlate with the
relative proportions of suitably thick and mature
units near reservoirs, given that most Cretaceoussourced oil appears to have escaped prior to trap
development. Maui family oils do appear to be primarily sourced by Rakopi Formation (late Cretaceous) coals. In the northern part of the Taranaki
Peninsula, where heat flows are highest, Mangahewa/Kaimiro Formation (Eocene) coals are the chief
sources of oils. Farther south, in the Kapuni and
Kupe South fields, Farewell Formation (Paleocene)
coals appear to be the main oil source rocks.
Biomarkers suggest that the onset of oil expulsion
from coals occurs at a maturity level corresponding
to a vitrinite reflectance of ca. 0.8% Ro, and may be
aided by the evolution of large volumes of carbon
dioxide. The terrestrial influence on Paleogene
source rocks diminishes to the north-northwest of
the basin and increasing marine contributions to
oils are observed. A late Paleocene marine black
shale is the source of oil in the Kora volcanic strucCopyright 1994. The American Association of Petroleum Geologists. All
rights reserved.
1Manuscript received July 27, 1993; revised manuscript received May 4,
1994; final acceptance May 27, 1994.
2Institute of Geological and Nuclear Sciences, P.O. Box 30368, Lower
Hutt, New Zealand.
3Industrial Research Ltd., Gracefield Research Centre, P.O. Box 31310,
Lower Hutt, New Zealand.
We wish to thank colleagues at IGNS, particularly Rick Allis and Rob
Funnell, and also Phil Armstrong, University of Utah, and Vanessa Killops for
help in the preparation of this article.

1560

ture. It is possible that shales interbedded with


coals, ref lecting periodic marine incursions of
coastal flood plains, also contribute to oil generation throughout much of the basin.
INTRODUCTION
The Taranaki Basin, which can be divided into
the Western Stable platform and the Eastern Mobile
belt, covers an area of some 100,000 km2 lying off
the west coast of the North Island of New Zealand
and extending under the post-Pliocene volcanic
deposits of the Taranaki Peninsula (Figure 1). It is
the primary petroleum-producing basin of New
Zealand, with an estimated 1678 106 bbl of recoverable petroleum (1213 106 bbl equivalent of gas,
210 106 bbl of condensate, 255 106 bbl of oil) in
the main fields (Figure 1 and Table 1). Oil accumulations are commonly accompanied by large volumes of gas and condensate, and all petroleum discoveries to date have been associated with
structural traps developed during the Neogene.
Reservoirs include Eocene coastal/shore sandstones, Oligocene to early Miocene fractured limestones, Miocene volcaniclastics, and MiocenePliocene turbiditic/fan sandstones (Bennett et al.,
1992). Potential reservoir capabilities are also
demonstrated by Late Cretaceous sandstones.
On the basis of organic content, hydrogen richness, and maturity constraints, the coals of the Paleogene Kapuni Group, and in particular the Late Cretaceous Pakawau Group, have been considered the
most likely petroleum source rocks (e.g., Cook,
1987; Johnston et al., 1990; Thrasher, 1992). Attention has been paid, therefore, to the organic geochemistry of these coaly sediments in order to correlate oils and their source rocks (for which biomarker
analysis is a powerful tool) and to determine maturation history and petroleum-expulsion characteristics.
Previous biomarker studies have mostly centered on
the ubiquitous hopanes and steranes, and only limited source information has been gained, mainly from
the relative abundance of the angiosperm-derived
compound 18(H)-oleanane (e.g., Cook, 1987;
Cook, 1988; Czochanska et al., 1988; Johnston et al.,
AAPG Bulletin, V. 78, No. 10 (October 1994), P. 15601585.

Killops et al.

(A)

1561

(B)

Figure 1The Taranaki Basin and its major petroleum fields, structural features, well locations, and heat-flow patterns. (A) Petroleum fields and main fault systems. (B) Well locations and heat-flow patterns. (Heat flow after Allis,
personal communication. Location of cross sections in Figures 10 and 11 also shown. Ari = Ariki-1, Awk = Awakino1, Fre = Fresne-1, Kai = Kaimiro-1 and 2, KDp = Kapuni Deep-1, Kpe = Kupe-1, KS4 = Kupe South-4, McK = McKee-1
and 3A, NP2 = New Plymouth-2, Tne = Tane-1, Tng = Tangaroa-1, Wai = Waihapa-1 and 2, Wit = Witiora-1; see Table 2
for other well abbreviations. Republic New Plymouth wells 1 and 4 lie close to New Plymouth-2.)

1991). More recently, the source-specific potential of


gymnosperm-derived diterpanes (Weston et al.,
1989) and various angiosperm-derived triterpanes
(Woolhouse et al., 1992) has been recognized and
partially applied in subsequent correlation studies
(Johnston et al., 1990; Collier and Johnston, 1991;
Johnston, 1992), but precise oil-to-source-rock correlation has still proven elusive. The study reported in
the following sections attempts to use source indicators to determine genetically related families of oils
and to identify their most likely source rocks. Basinwide oil generation and expulsion is then considered in the light of the distribution and maturity of
potential source rock units.
The geology of the Taranaki Basin is complex
(e.g., King, 1990; Palmer and Bulte, 1991; King

and Thrasher, 1992), and the following summarizes the depositional environment of potential
source rocks and their spatial and temporal relationships with traps. Extension of the Paleozoic to
middle Cretaceous basement began ca. 80 m.y.a.
and led to the development of north-northeast
trending subbasins (up to 100 km long by 30 km
wide), in which up to ca. 3000 m of Late Cretaceous sediments accumulated. These are the terrestrial sediments of the Pakawau Group, which
are thicker to the west of the basin and include the
coals of the Rakopi Formation (Figure 2). Toward
the end of the Cretaceous, a rapid, southeast
trending, marine transgression resulted in deposition of the mudstones and sandstones of the North
Cape Formation (Figure 2).

1562

Oil Generation in the Taranaki Basin

Table 1. Proven, Recoverable, Petroleum Reserves in the Taranaki Basin*


Field
or Well

Reservoir
Formation

Kora
Mangahewa
Moturoa
Kaimiro

Mohakatino
Mangahewa
Matemateaonga
Mt. Messenger
McKee
Mt. Messenger
McKee/Mangahewa
McKee
McKee
Otaraoa
Otaraoa
Tikorangi
Kaimiro
Tikorangi
Mangahewa
Farewell
Farewell
Mangahewa/Kaimiro
Farewell
Moki
Mangahewa

Ngatoro
Stratford
Urenui-1
McKee
Tariki
Ahuroa
Waihapa
Ngaere
Kapuni
Toru
Kupe South
Maui
Moki
Maui-4

Oil
(106 bbl)

Condensate
(106 bbl)

Gas
(106 bbl
oil equivalent)

<1
na

38

21

7
128

10
36
7

na
1

2
1

<1

55

151

na
6
<1
4
6
20
10
5
4
3
1
196
20
144
792

*1 bbl oil equivalent = 150 m3 gas. See Figure 1 for field locations. na = data not available.

Tectonic activity waned during the Paleocene,


resulting in continued marine incursion. Extensional faulting ceased by the end of the Paleocene,
subsequent passive-margin cooling and subsidence
led to a southwest-directed marine transgression
over coastal plains, and sediment supply diminished. During the tectonic quiescence of the Paleogene, the terrestrial sediments of the Kapuni
Group were deposited, dominated by sandstones
and lower flood plain coal measures with some
shales. Four regressive cycles have been recognized in which coastal/marginal marine sandstones pass upward into coal-bearing intervals
(e.g., Johnston et al., 1990). The Kapuni Group is
extensive in the southeast of the basin (Farewell,
Kaimiro, Mangahewa, and McKee formations, Figure 2), whereas to the north and northwest a
marine equivalent can be recognized in the darkbrown to gray, partly carbonaceous and micaceous, marine siltstones and mudstones of the Turi
Formation (Figure 2). In the northwest, the
Eocene/Oligocene Tangaroa Formation represents
a sandy submarine-fan complex. The transgression
reached a maximum in the early Oligocene, covering the entire basin and marking the culmination
of passive-margin subsidence. The terrigenous supply of sediment became limited. However, rapid
subsidence recommenced in the mid-Oligocene.
Oligocene sedimentation is characterized by lime-

stones of the Tikorangi Formation and mudstones,


siltstones, and sandstones of the Otaraoa Formation (Figure 2).
In the early Miocene, uplift of the hinterland and
overthrusting along the eastern margin led to
increased sediment supply and a regressive sedimentation pattern, which has persisted to the present. The Manganui Formation represents mudstones deposited off the shelf edge, whereas the
Moki Formation comprises slope-fan sandstones
(Figure 2). The Mohakatino Formation in the northern part of the basin consists mainly of stratified
volcaniclastic deposits of the middle to late
Miocene, derived from andesitic volcanoes associated with subduction of the Pacific plate. Along the
northeast margin of the basin, turbiditic sandstones
of the Mount Messenger Formation were deposited, overlain by finer slope sediments of the Urenui
Formation (Figure 2). At this time, compression in
the south of the basin initiated the inversion of
existing structures. Renewed uplift of the eastern
hinterland increased the sediment supply during
the PliocenePleistocene, with the accumulation of
shelf conglomerates, sandstones, and mudstones of
the Matemateaonga Formation in the east and the
Giant Foresets Formation (prograding shelf mudstones and minor sandstones) farther west, the latter representing most of the sediment on the Western Stable platform (Figure 2).

Killops et al.

1563

Figure 2Generalized stratigraphy of


the Taranaki Basin (after Bennett et al.,
1992), showing main oil/condensate
accumulations. (KS = Kupe South field
oils; MK = oils from McKee-3A, Pouri1A, Pukemai-2A, Stratford-1, Toetoe-1
and Urenui-1; Mu = oils from Maui-1
and Maui-3; NP = oils from Moturoa-2,
Republic New Plymouth-1, Republic
New Plymouth-4 and Taranaki-5; see
Table 2 and Figure 1 legend for other
well abbreviations.)

MATERIALS AND METHODS


The locations of wells from which oils/condensates and organic-rich sediment samples were
taken are shown in Figure 1. Wells and formations
from which oil/condensate (DST or production)
samples were taken are listed in Table 2. Samples
from Kapuni-2, Maui-3, Toru-1, Urenui-1, and Waihapa-1A are classified as condensates, whereas samples from Okoki-1, Pukearuhe-1, and Tangaroa-1 are
oil shows. Additional biomarker analyses were carried out on condensates and oils from Kupe South
wells 25; a condensate from Maui-1; oils from
McKee wells 1, 2A, and 4; and oil from Toetoe-2. All
these samples were found to be chemically almost
identical to other samples from the fields concerned and so are not considered further here.
Cuttings were selected from the following formations: Farewell (Kapuni Deep-1, Kupe-1, Kupe
South-1, Kupe South-4, and Maui-4); Kaimiro
(Kaimiro-1, Kapuni Deep-1, Maui-4, and Waihapa1); Mangahewa (Kaimiro-1, Kapuni Deep-1, McKee1, Maui-4, Urenui-1, and Waihapa-1); North Cape
(Kupe South-4, Maui-4, and Tane-1); Rakopi (Fresne1, Maui-4, and Tane-1); and Turi (Ariki-1, Awakino1, Kaimiro-1, and Witiora-1). For some samples, a
coal fraction was separated from shale by a rapid

flotation method using a carbon tetrachloride/hexane mixture with a density of 140 kgm 3. After
being milled to a fine powder, sediment samples
were solvent-extracted using either Soxhlet extraction or ultrasound. Saturate fractions were isolated
from oils and sediment extracts by TLC on silica gel
with hexane eluant. Where necessary, n-alkanes
were removed from saturate fractions by 5 molecular sieve or urea adduction prior to gas chromatographymass spectrometry (GCMS) analysis of
biomarkers using columns coated with 5% phenylmethylsilicone stationary phase and monitoring of
selected ions. Relative biomarker abundance was
calculated from peak heights in m/z 123 (sesquiterpanes and diterpanes), m/z 191 (triterpanes,
including hopanes), and m/z 217 (steranes) mass
chromatograms. From these data, the following
indices and ratios were obtained (see Figure 4 legend for key to compound abbreviations).
Angiosperm/gymnosperm index:
AGI =

m/z191 (O + dO + dL + dU)
m/z123 (L + R + NT + B + 18NIP + 19NIP + IP + P

m/z123 (IP)
m/z191 (IP)

NP1
NP4
Trn
Mot
Man
Km1
Km2
Nga
Str
Ure
Tuh
Pou
Pkm
MK3
Toe
Tik
Ahu
Wp1
Wp2
Kap
Tou
KS1
Mu3
Mu1
Mki
Mu4

New Plymouth-1
New Plymouth-4
Taranaki-5
Moturoa-2
Mangahewa-1
Kaimiro-1
Kaimiro-2
Ngatoro-1
Stratford-1
Urenui-1
Tuhua-1
Pouri-1A
Pukemai-2A
McKee-3A
Toetoe-1A
Tariki-1A
Ahuroa-1
Waihapa-1A
Waihapa-2
Kapuni-2
Toru-1
Kupe South-1
Maui-3
Maui-1
Moki-1
Maui-4

Otaraoa
Otaraoa
Kaimiro
Tikorangi

McKee overthrust
McKee overthrust
McKee overthrust
McKee overthrust
McKee overthrust
McKee overthrust

Mangahewa
McKee
Mt. Messenger
Mt. Messenger
McKee

Matemateaonga
Matemateaonga
Matemateaonga
Matemateaonga

Mohakatino
Turi
Basement**
Mangahewa

Maui
Maui
Moki
Maui-4

Mangahewa
Mangahewa
Moki
Mangahewa

Kapuni
Mangahewa
Toru
Farewell
Kupe South Farewell

Tariki
Ahuroa
Waihapa
Waihapa

Urenui
McKee
McKee
McKee
McKee
McKee

Mangahewa
Kaimiro
Kaimiro
Ngatoro
Stratford

Moturoa
Moturoa
Moturoa
Moturoa

Kora

Reservoir
Formation

48
44
37
44

23

48

47

48
38
37

39

41
42

34
32
31
37

35

30

API
gravity
()

0.08
0.06
0.11
0.30

0.06
0.02
0.12

0.15
0.10
0.03
0.08

0.16
0.10
0.13
0.08
0.24

0.07
0.10
0.05
0.06
0.10

0.09
0.08
0.07
0.08

0.27

0.18

S
(%)

87
94
81
89

94
95
84

77

92
93
94
93
96
89

95
85

93

93
90
92
91

94

95
75

HC
(%)

79
93
77
84

81
87
92

76

86
87
86
82
87
85

83
76

86

75
82
75
76

69

76
40

Sats.
(% of
HC)

(27.1)
27.8
27.9
28.1

28.4

28.4

28.8
28.5

28.4
28.7
28.7
28.6
28.8
28.8

28.2
28.3

28.8

(28.2)
(28.2)
28.8
28.8

23.6

28.2

13C
Sats.
()

(26.2)
26.2
26.1
26.2

26.4

26.4

26.3
26.0

26.2
26.3
26.5
26.4
26.4
26.6

25.9
26.1

26.4

(26.5)
(26.4)
26.8
26.6

22.0

26.3

13C
Aroms.
()

(26.5)
27.3
27.3
27.4

27.9

27.5

28.1
27.7
28.0

28.0
28.0
28.3
28.0
28.2
28.4

27.5
27.6
27.4
27.4
28.3

(27.2)
(27.5)
28.1
28.0

22.9
23.7
27.5

13C
Total
()

+13.2
+9.6
+13.7

+15.6

+16.3

+8.1
+13.8
+11.3
+15.1

+9.4
+13.3
+10.6
+13.3
+12.4

+14.1
+11.4
+11.8
+8.6

+3.0
+6.4
+12.7
+4.7

8.0

+6.0

34S
Total
()

1.00

0.99

1.00

1.00

1.00

0.98

V
V + Ni

9.0
6.1
8.3
8.3

8.2
7.2
6.3

7.6
7.4
7.9
8.0

9.0
8.2
7.4
7.2
7.5
8.3

6.0
6.6
6.5
4.5
9.2

7.0
7.2
5.9
7.0

2.5
2.7
5.7
5.5

Pr
Ph

0.9
1.3
1.2
1.2

1.1
1.1
1.1

1.2
1.2
1.1

1.1
1.1
1.2
1.2
1.1
1.2

1.1
1.1

1.1
1.2
1.1
1.2

1.0
1.1
1.2
1.1

CPI

*Data for Tangaroa-1 after Gibbons et al., 1981; for Okoki-1 after Analabs, 1989; and for Toru-1 after Analabs, 1991. Other carbon isotopic data (relative to PDB) mainly after Weston et al., 1988b.
Accuracy of 13C data in parentheses (after Hirner and Lyon, 1989) is doubtful, probably related to incomplete solvent removal from saturate fractions. Sulfur isotopic data (relative to CDT) after Hirner and
Robinson, 1989, and Robinson, personal communication; and trace-metal data after Frankenberger et al., 1994. Pr/Ph = pristane:phytane; CPI = carbon preference index (Killops and Killops, 1993); Sats. =
saturates; Aroms. = aromatics; dashes = data not available. The two New Plymouth oils are from wells drilled by the Republic Oil Co.
**In Pukearuhe-1, Paleogene and Cretaceous strata are absent and oil show was in fracture zone associated with intrabasement thrust in late Triassic to early Jurassic sediments.

Kor
Tng
Pkh
Oko

Kora-1
Tangaroa-1
Pukearuhe-1
Okoki-1

Well
Abbr. Field

Table 2. Geochemical Characteristics of Whole-Oils*

1564
Oil Generation in the Taranaki Basin

Killops et al.

(The m/z 123/1 91 response factor for isopimarane


is used to obtain an equivalent m/z 191 response
for all the diterpanes, most of which do not yield
appreciable m/z 191 responses, for comparison
with the m/z 191 responses of the triterpanes.)
Terrestrial/marine index:
TMI =

m/z217 (29DS + 29DR)


m/z217
(27DS + 27DR + 30DS + 30DR)

Hopane:sterane ratio:
H/S =

m/z191 (dH + 30 + 30 )
m/z217 (27 + 29 + 30)

[where m/z 217 response includes C27, C29, and C30


5(H)-steranes and 13(H),17(H)-diasteranes]
Tricylic:tetracyclic terpane ratio:
TTR =

m/z191 (23T)
m/z191 (dH)

Most of the biomarker data were obtained from


two sample sets analyzed on different GCMS systems (see Table 3). To facilitate comparison of
these two data sets, some oils were analyzed on
both systems to permit compensation for variations
in the relative responses of components between
GCMS systems.
OIL-GEOCHEMISTRY INFERENCES FOR
SOURCE AND DEPOSITIONAL ENVIRONMENT
The bulk geochemical and biomarker characteristics of the oils/condensates (hereafter collectively
referred to as oils) analyzed are listed in Tables 2
and 3, well locations are shown in Figure 1, and the
formations associated with reservoirs/shows are
shown in Figure 2. Relative terrestrial and marine
contributions to the source rocks of the oils can be
evaluated from a plot of 13C for aromatic vs. saturated hydrocarbon fractions, as shown in Figure 3.
Oils tend to plot on or near the terrestrial or marine
lines in Figure 3, depending on the origins of their
source kerogens (Sofer, 1984). Most Taranaki Basin
oils appear to be predominantly terrestrial in origin, with limited variation in carbon isotopic composition (Table 2). This is consistent with the waxy
nature of these oils, high hopane:sterane ratios, and
high relative proportions of C29 steranes (Table 3).
C29 sterols are the dominant phytosterols in terrestrial higher plants but are generally less abundant in
plankton, and so dominant C 29 steranes usually

1565

indicate significant contributions from terrestrial


higher plants, although care is required in applying
this ratio in the absence of information on stereochemistry at C-24 (Killops and Killops, 1993, and
references therein). This relationship for Taranaki
Basin oils can be expressed by the ratio of C29 to
C27 and C30 regular steranes, the terrestrial/marine
index (TMI), because the relative abundance of C28
steranes is reasonably constant (Table 3). The
hopane:sterane ratio (H/S, Table 3) provides a measure of bacterial activity, particularly that associated
with the degradation of the lignified tissues of higher plants resulting in the concentration of lipid-rich
structures, such as leaf cuticles and waxes and bacterial membranes. It is not unexpected, then, that
the highest hopane:sterane ratios are recorded for
the most waxy oils, from Kupe South-1, McKee-3A,
Pouri-1A, and Stratford-1, and that a reasonable positive correlation is found between H/S and TMI
(correlation coefficient, r = 0.71).
A slight odd-over-even predominance (OEP) is
present among n-alkanes in the C23C33 range for
the terrestrial oils (CPI ca. 1.1, Table 2), characteristic of a higher plant epicuticular wax contribution
(Baker, 1982). Typical biomarker distributions are
shown in Figure 4. Angiosperm contributions are
indicated by the presence of oleanane (O) and the
C24 ring-A degraded counterparts of oleanane (dO),
lupane (dL), and ursane (dU) in the m/z 191 mass
chromatogram (which probably derive from microbial alteration of the pentacyclic analogues; Woolhouse et al., 1992, and references therein). The five
C30 pentacyclic triterpenoids marked by asterisks
(Woolhouse et al., 1992) are thought to be
angiosperm-derived because their abundances in
the oils vary in a manner broadly similar to that for
oleanane. All oils contain 18(H)-30-norneohopane
(29N) and 17(H)-15-methyl-27-norhopane (or
diahopane, 30D), which have also been observed
in various terrestrial oils from around the world,
usually accompanied by 18(H)-oleanane (Killops
and Howell, 1991, and references therein). The
peak immediately to the left of 27 in the m/z 191
mass chromatogram in Figure 4 is an unidentified
triterpane that co-elutes with trans,trans,transbicadinane (BC), a compound believed to derive
from cadinene polymers in angiosperms of the
dipterocarp family (van Aarssen et al., 1992).
Because of the co-elution, the relative abundance
of bicadinane cannot be ascertained from m/z 191
mass chromatograms alone, but traces of this compound do appear to be present in most oils from
characteristic m/z 369 and 412 responses.
The angiosperm-derived sesquiterpenoids eudesmane and bisabolane (E and BA, respectively, m/z
123 mass chromatogram, Figure 4) are present in
low levels in all the oils analyzed (ca. 5% relative to
drimane, D). The ratio of 8(H)-homodrimane

1566

Oil Generation in the Taranaki Basin

Table 3. Aliphatic Biomarker Ratios of Oils (All Values as Percentages)*


HD
HD+D

Sesqui-/Diterpanes
RD1+RD2
IP
RD1+RD2+D
DT

HB
DT

27:28:29:30

29D
29

Steranes
S

S+R
+

TMI
(%)

Kora-1
Tangaroa-1
Pukearuhe-1
Okoki-1

70

66
73

69

64
76

28
24

24

18
26

30:10:19:41
34:13:21:32
28:20:34:18
35:26:39:0

31
34
57
42

53
59
58
45

58
57
48
49

36
32
74
111

New Plymouth-1
New Plymouth-4
Taranaki-5
Moturoa-2

63
71
72

53
64
69

24
16
15

33

21:25:54:0
25:21:54:0
23:22:55:0
34:24:42:0

53
61
51
64

53
61
61
55

48
44
46
48

257
216
239
124

Mangahewa-1
Kaimiro-1
Kaimiro-2
Ngatoro-1
Stratford-1

63
66
75
77
82

64
57
58
56
45

61
28
35
46
27

8
21
13
9
13

42:24:34:0
33:22:45:0
26:23:51:0
23:20:57:0
23:22:55:0

71
63
57
48
47

60
59
62
59
49

51
51
47
46
44

81
136
196
252
239

Urenui-1
Tuhua-1
Pouri-1A
Pukemai-2A
McKee-3A
Toetoe-1A

66
71
73
66
77
71

60
59
58
64
55
56

31
15
34
21
17
9

12

12
13
10
19

14:21:65:0
11:15:74:0
11:17:72:0
14:18:68:0
12:18:70:0
18:19:63:0

62
54
46
46
39
59

52
53
55
50
46
49

45
41
43
40
44
49

464
672
655
486
583
350

Tariki-1A
Ahuroa-1
Waihapa-1A
Waihapa-2

71
62
56
73

64
50
74
44

22
26
54
36

22
21
6
14

12:19:69:0
18:19:63:0
24:26:50:0
12:16:72:0

53
62
45
44

44
53
47
60

44
47
46
45

575
350
208
615

Kapuni-2
Toru-1
Kupe South-1

64
52
65

43
44
44

51
50
46

12
12
9

12:23:65:0
12:20:68:0
10:22:68:0

56
37
47

53
46
46

47
49
43

542
567
680

Maui-3
Maui-1
Moki-1
Maui-4

69
57
60
61

66
65
58
61

62
55
60
62

3
2
4
3

15:19:59:7
21:21:54:4
23:21:52:4
19:20:57:4

51
61
66
55

51
53
55
50

48
50
53
49

268
216
193
248

*Sesqui-/diterpanes: DT = sum of diterpanes labeled in Figure 4. Steranes: 29D/29 = ratio of 13(H),17(H),20S/R-diasteranes to sum of C29 5(H)steranes and 13(H),17(H)-diasteranes; S/(S + R) = ratio of 20S to 20R for 5(H),14(H),17(H)-24-ethylcholestanes; /( + ) = ratio of
5(H),14(H),17(H)- to 5(H),14(H),17(H)-24-ethylcholestane (20S + 20R); TMI, see Materials and Methods section. Triterpanes: /( + ) = ratio of
17(H)-hopane to 17(H)-moretane; S/(S + R) = ratio of 22S to 22R for 17(H)-homohopane; AGI and H/S, see Materials and Methods section. For
identification of all other components, see Figure 4 legend. Data for Tangaroa-1 after Gibbons et al., 1981; for Okoki-1 after Analabs, 1989; and for Toru-1
after Analabs, 1991. nd = 28B not detected.

(HD) to 8(H)-drimane (D) is similar for most oils


[HD/(HD + D), Table 3], a possible exception being
Stratford-1. It is probable, therefore, that D and HD
are primarily derived from hopanoidal precursors
rather than there being a significant contribution to
drimane from higher-plant-derived drimenol.
All of the labeled diterpanes in the m/z 123 mass
chromatogram in Figure 4 except HB (which is
probably of bacterial origin) are believed to be
derived from gymnosperms (especially podocarps
and araucarians; Noble et al., 1986; Weston et al.,
1989). Distributions of diterpanes other than isopi-

marane (IP) and the C 21 bicyclane (HB) are quite


uniform and so only relative abundances of IP and
HB are given in Table 3. Abietane (A) is never more
than a very minor diterpenoid in Taranaki oils (1%
relative to total diterpanes).
The 13C plot in Figure 3 indicates that there is
only one primarily marine oil, that from Kora-1.
The marine signature in this oil is characterized by
a low TMI value (Table 3) and unusually abundant
C30 regular steranes (Figure 5, m/z 217 mass chromatogram), believed to be derived from chrysophyte algae (Moldowan et al., 1990). However,

Killops et al.

1567

Table 3. Continued.
Triterpanes
30D
30

S
S+R

27N
27N+27

dL
dL+dO

89
89
90
88

55
59
58
60

54
49
42
44

18

13
14

1
2
2
9

13
6
27
15

90
90
92
91

60
60
59
60

52
42
58
53

9
15
4
12

8
2
1
2

90
89
87
86
87

61
60
61
61
60

35
21
20
16
10

21
29
30
41
65

88
89
90
88
88
90

62
59
60
59
59
58

24
18
18
21
18
23

89
91
82
90

59
60
60
60

88
85
87
87
90
88
88

28B
30

AGI
(%)

23T
dH

H
S

104
96
129
129

61

22
80

100
100
79
74

31
45
58
94

30
26
31
27

108
114
87
132

52
66
39

16
15
21

278
155
154

1
2
2
1
1

19
9
11
9
6

162
180
145
162
192

22
46
51
41
55

26
33
15
14
15

90
164
298
332
287

52
41
43
28
40
46

nd
nd
nd
nd
nd
1

14
15
12
12
12
14

187
156
148
165
188
175

75
188
112
30
388
68

25
8
6
14
14
13

117
196
354
247
343
157

14
23
28
16

52
38
6
48

2
1
nd
<1

8
15
33
6

172
141
222
134

148
74
13
35

20
15
21
5

160
146
162
509

60
61
58

19
19
17

38
38

2
6
1

13
14
11

167
156
167

19
14
19

11
13
12

201
142
335

60
58
60
54

29
48
35
37

15
16
34
43

2
6
5
6

13
19
27
23

139
131
164
154

3
2
10
5

32
31
38
21

161
67
113
105

Kora-1 is not entirely marine sourced, because it


contains oleanoid biomarkers of angiosperm origin. Similar biomarker characteristics are also
exhibited by the oil show in late Paleocene strata in
Tangaroa-1 (Table 3), suggesting a common source
with Kora-1 oil. Carbon isotopic data (Figure 3),
TMI value, and C30 sterane content (Table 3) suggest that there is a smaller contribution from a
marine, Kora-type source to the oil show from
Pukearuhe-1. Traces of C 30 regular steranes were
also observed in Maui-1 and Maui-4 oils (from
molecular ion and m/z 217 responses), and possibly also in Maui-3 and Moki-1 (from m/z 217
responses only), suggesting a minor marine contribution, which is also supported by carbon isotopic
data (Figure 3) and intermediate TMI and low H/S
values (Table 3). In all the oils exhibiting identifi-

29+31
30

able marine contributions, the level of tricyclic terpanes relative to hopanoidal terpanes is higher
(23T/dH, Table 3), which is consistent with the tricylics being derived from bacteria dwelling in
saline environments (De Grande et al., 1993). As
might be expected, therefore, TMI and 23T/dH values exhibit a negative correlation coefficient (r =
0.61). On the basis of TMI, 23T/dH, and H/S values, there appears to be a marine contribution to
the source rocks of the oils from Kaimiro-1, Moturoa-2, and Okoki-1, although C 30 regular steranes
were not detected. A marine influence in the oils
also appears to correlate with lower (29 +
31)/30 hopane ratios (Table 3), which is consistent with the greater importance of the C31 and
C 29 components in lignites (e.g., Brassell et al.,
1986; Lu and Kaplan, 1992). C30 4-methylsteranes

1568

Oil Generation in the Taranaki Basin

Figure 3Plot of 13C values for


aromatic vs. saturated hydrocarbon fractions from Taranaki
Basin oils. (Data after Weston et
al., 1988b, and Reed, 1992; terrestrial and marine classification
after Sofer, 1984. See Table 2 for
well abbreviations.)

have been detected in all oils for which suitable


GCMS data (m/z 231 responses) are available (Kora1, Kupe South-1, Maui-1, Maui-4, Okoki-1, and
Pukearuhe-1). For most of these oils, a marine
dinoflagellate origin for the 4-methylsteranes is likely (Summons et al., 1987, and references therein).
All oils have low sulfur contents (0.3%; Table 2)
and low levels of nickel and vanadium (<2 ppb and
<650 ppb, respectively; Frankenberger et al.,
1994), probably reflecting low levels of asphaltenes/resins (mostly <20%), with which most sulfur
and trace metals are usually associated. Oils for
which nickel and vanadium concentrations are
available (Frankenberger et al., 1994) exhibit V/(V
+ Ni) ratios approaching unity, which, together
with sulfur contents of 0.3% or less (Table 2), suggest limited sulfate-reducing bacterial activity
(Lewan, 1984). The 34S values for most oils (Table
2) fall within the range previously attributed to
restricted marine or high-rate sedimentation environments, in which sulfate supply and hence sulfate-reducer activity are limited by the rate at which
bacterial sulfide oxidation occurs (Hirner and
Robinson, 1989). Such indications can be reconciled with predominantly terrestrial depositional
environments undergoing minor marine incursions. Because sulfur levels are low in terrestrial
and freshwater plants, only a small marine sulfate
contribution is required to yield the observed low
sulfur levels that exhibit an isotopic signature closer to ambient marine (+19 < 34S < +21) than freshwater (+1 < 34S < +7) sulfate (Hirner and Robinson, 1989). Periodic marine innundations of coastal
flood plains permit sulfate to percolate down into
previously deposited freshwater peats (Suggate,
1959), resulting in stimulation of sulfate-reducer

activity and subsequent diagenetic incorporation of


sulfide into the organic matrix of peat. Because little microbial isotopic fractionation would be
expected from such a limited sulfate supply, this
mechanism explains the high positive 34S signature (approaching that of ambient seawater) of
coals from Taranaki wells (Table 4). Shales interbedded with the coals exhibit the greater isotopic fractionation (i.e., negative shift in 34S values, Table 4)
and sulfur levels expected for a less restricted sulfate supply. If source rock coals and shales throughout the Taranaki Basin exhibit these sulfur isotopic
patterns, it would seem that coals are the main
source of most of the oils for which sulfur isotopic
data are available (Table 2), but that there is a significant marine shale contribution to Moturoa-family
oils and the oil show in Pukearuhe-1. The sulfur isotopic signature for Kora-1 oil is consistent with an
open-marine depositional environment.
Although pristane:phytane ratios (Pr/Ph, Table 2)
can be significantly source-dependent, it appears
that redox conditions are reflected in the values
reported for Taranaki oils (Killops and Killops, 1993,
and references therein). Values for the predominantly terrestrially-sourced oils lie in the range ca. 79,
suggesting that associated depositional environments were not entirely anoxic. The chiefly marinesourced oils from Kora-1 and Tangaroa-1 exhibit
lower Pr/Ph values (<3), which probably reflect less
oxygenated conditions. In peat swamp environments, it can be envisaged that the sheer quantity of
higher plant organic detritus overwhelms the capacity of the decomposer community to degrade it fully,
allowing preservation of organic-rich sediments
under relatively oxidizing conditions and the concentration of more resistant, hydrogen-rich compo-

Killops et al.

1569

Figure 4Typical distributions of terpanes and steranes in terrestrially sourced oils of the Taranaki Basin. [In
sesquiterpanes: D = 8(H)-drimane; HD = 8(H)-homodrimane; RD1 and RD2 = rearranged drimanes; E = 4(H)eudesmane; BA = bisabolane. In diterpanes: L = 8(H)-labdane; 18NIP and 19NIP = 4(H)-18- and 4(H)-19-norisopimarane; R = rimuane; NT = 17-nortetracyclane (C19); B = ent-beyerane; IP = isopimarane; HB = homobicyclane (C21);
P = 16(H)-phyllocladane; K = ent-16(H)-kaurane; A = abietane. In triterpanes: numbers correspond to carbon
numbers; T = tricyclic terpane; dO, dL, and dU = 10(H)-des-A-oleanane, -lupane, -ursane; dH = 18(H)-des-Ehopane; 27N = 18(H)-22,29,30-trisnorneohopane; +BC = unidentified triterpane + trans,trans,trans-bicadinane;
27 = 17(H)-22,29,30-trisnorhopane; 28B = 17(H),18(H)-28,30-bisnorhopane; C 29-C 31 hopanes are labeled
according to stereochemistry at C-17 and C-21, and also at C-22 for 31; 29N = 18(H)-30-norneohopane; 30D =
17(H)-15-methyl-27-norhopane; O = 18(H)-oleanane; * = C30 pentacyclic triterpanes. In steranes: numbers correspond to carbon numbers; regular (4-desmethyl) sterane stereochemistry at C-14, C-17, and C-20 is given, all components have 5(H) configuration; diasteranes are indicated by D and stereochemistry at C-13, C-17, and C-20.]

nents. In fully marine environments, preservation of


sufficient phytoplankton-derived sedimentary organic matter to generate petroleum is generally associated with high primary productivity and dysaerobia or
anoxia, resulting either from the impingement of an
open-ocean, oxygen-minimum layer on the shelf/
slope or from oxygen depletion during degradation
of large amounts of phytoplankton detritus in an
area with restricted inflow of fresh, oxygenated bot-

tom water. Most Taranaki oils seem to contain small


amounts of 17(H),18(H),21(H)-28,30-bisnorhopane (28B/30, Table 3). This compound
appears to be characteristic of oxygen-deficient,
restricted marine, depositional environments (Mello
et al., 1988; Burwood et al., 1992), and again suggests that there is a small marine contribution to the
source rocks of even those oils exhibiting the most
terrestrial characteristics.

1570

Oil Generation in the Taranaki Basin

Figure 5Mass chromatograms showing triterpane and


sterane distributions in Kora-1 oil and bitumen from
Waipawa Black Shale equivalent in Ariki-1 well. [Triterpanes: 27 = 17(H)-21,29,30-trisnorhopane, 29 =
17(H)-30-normoretane, 31 = 17(H)-31-homomoretane. Steranes: 21 = diginane, 22 = 20-methyldiginane,
30DS/R = 20S/R epimers of 13(H),17(H)-24-n-propyldiacholestane, 30R = 20R-5(H),14(H),17(H)-24-npropylcholestane, 30 = 5(H),14(H),17(H)-24-npropylcholestane. See Figure 4 legend for other
component identifications.]

A final observation about depositional environment can be made regarding the high levels of rearranged relative to regular steranes in the terrestrial
oils compared with those in the predominantly
marine-sourced oils (29D/29, Table 3). Rearrangement of steroids occurs during diagenesis following
dehydration of sterols to sterenes, and is catalyzed
by clays (Killops and Killops, 1993, and references
therein). However, the extent to which clay-catalysis contributes to steroid rearrangement in low-ash
New Zealand coals is not certain. Dominance of
diasteranes over regular steranes may primarily
result from the greater resistance of diasteranes
toward biodegradation during diagenesis. The generally low levels of steroids probably reflects bacterial degradation of mesophyll, with which most of
the higher-plant phytosterols are associated. There
appears to be some correlation between the relative abundances of C24 tetracyclanes (dO, dL, dU,
and dH) and diasteranes, which could be related to
microbial activity.
A clay source has been suggested for several
trace elements, such as aluminum, in Taranaki oils
(Frankenberger et al., 1994). Although it is possible
that trace elements in clay minerals are incorporated into the organic macromolecular structure of
the source rock during diagenesis and are liberated
as an integral part of the asphaltenes fraction during catagenesis, it would seem unlikely that colloidal clay particles would undergo migration from
the source rock with the liquid hydrocarbon phase
(Frankenberger et al., 1994) on the basis of size
restrictions on migration. It is more likely that clay
minerals from reservoir formations (e.g., Hill and
Collen, 1978) may become dispersed in the oils.
Whatever the origin of the clay minerals, it is not
directly linked to the sources of the organic material from which the oils are generated, and so the
proposition that all Taranaki oils apart from Maui-1
belong to the same genetic family, based on tracemetal content (Frankenberger et al., 1994), is not
necessarily valid. Indeed, various plots of sourcerelated biomarker parameters suggest that four
main families of oils can be recognized: Kapuni
(Kapuni-2, Kupe South-1, and Toru-1); McKee

Killops et al.

1571

Table 4. Sulfur Stable Isotopic Compositions for Selected Coal and Shale Samples*
Depth
(m)

Formation

Sediment

%S

Kerogen
34S

McKee-1

36623672
36623672

Mangahewa
Mangahewa

Shale
Coal

5.5
2.1

+10.0
+19.4

Maui-1

31033106
31063109
33623365
33653367

Kaimiro
Kaimiro
Farewell
Farewell

Shale
Coal
Coal
Shale

3.4
0.1
2.8
6.1

+5.7
+3.3
+3.5
7.0

Maui-4

20912118
32463252
38223828
38283837

Mangahewa
Rakopi
Rakopi
Rakopi

Coal
Coal
Coal
Shale

0.3
0.6
0.7
8.0

+11.9
+19.5
+12.0
+9.1

4.3
4.3
4.6
3.4

Fresne-1

24862501
24862501

Rakopi
Rakopi

Shale
Coal

5.4
0.9

+2.7
+16.4

2.4
8.4

Castlepoint**

Outcrop

Turi (Waipawa)

Shale

2.1

6.1

0.4

Well

Approx.
TMI

*After Hirner and Robinson, 1989.


**Castlepoint sample of Waipawa Black Shale is from East Coast Basin.

(McKee-3A, Pouri-1A, Pukemai-2A, Toetoe-1A,


Tuhua-1, and Urenui-1); Maui (Maui-1, Maui-3, Maui4, and Moki-1) and Moturoa (Moturoa-2, New Plymouth-1, New Plymouth-4, and Taranaki-5). In addition, oils from Kaimiro, Ngatoro, and Stratford
fields share some similarities, and oils from Ahuroa1 and Tariki-1A exhibit some affinities with the
McKee family. These associations can be seen in
three plots based on terpane distributions from
m/z 123 and 191 mass chromatograms in Figure 6.
The enhanced levels of tricyclic terpanes in the
most strongly marine-influenced oils from Kora-1,
Okoki-1, Pukearuhe-1, and Tangaroa-1 can be seen
in the plot of 27N/(27N + 27) vs. 23T/(23T + dH)
(Figure 6). The ratio 27N/(27N + 27) is often used
as a maturity indicator, but it is also significantly
affected by variations in sources of organic matter.
Whether the genetic characteristics are sufficiently
distinct for specific source rock intervals to be
identified is considered in the following section.
OIL-TO-SOURCE ROCK CORRELATION
The selection of potential source rock samples
for correlation studies was based on measurements
of genetic potential and estimates of horizons likely
to have entered the oil window. Maturity considerations suggest that Pakawau and Kapuni Group coals
and shales are the most probable source rocks.
Genetic potential is summarized in Table 5 for
some formations that have previously been systematically screened by TOC and Rock-Eval analysis
(Analabs, 1984).

Paleogene and Late Cretaceous coals appear to be


reasonable candidates for the source rocks of the
more terrestrial oils, and can be seen to possess
mixed oil and gas potential (Table 5). The Rakopi
Formation (Figure 2) is widespread and contains
abundant coal, on the basis of seismic reflection
characteristics (Thrasher, 1992). However, the distribution of coal-rich units in Paleogene strata reflects
the general extent of marine transgressional cycles,
with progressively terrestrial-dominated sediments
being deposited toward the east-southeast. The
Eocene Mangahewa Formation (Figure 2) is particularly coal-rich and is found throughout the onshore
part of the basin and in the Maui-4 area. It is significantly thickened within the northern part of the
Tarata thrust zone (Figure 1). The Kaimiro Formation
is coal-rich within the Tarata thrust zone, but westward around New Plymouth (Figure 1) it is appreciably marine influenced and appears to have poor oil
potential. The Paleocene Farewell Formation generally contains less coaly material and exhibits varying
marine influence. In the Maui-4 area, it is predominantly terrestrial with oil potential, but in the Kupe
Field, only the lower half of the unit contains carbonaceous and coaly material in significant quantity
(Puponga Member), whereas the upper half is organic poor and marine influenced. The marine, Paleogene, Turi Formation predominates to the west and
north, and it is generally not particularly carbonaceous. Within the Turi Formation in the northwest
region of the basin there is an organic-rich shale
equivalent to the late Paleocene Waipawa Black
Shale, which is found in several other New Zealand
basins, such as the East Coast (Weston et al., 1988a),

1572

Oil Generation in the Taranaki Basin

are some broad similarities in biomarker distributions. For example, 10(H)-des-A-lupane is relatively abundant in McKee-family oils (Table 3) and in
Mangahewa Formation samples. The lack of perfect
matches is not surprising in view of the rapid
changes in flood plain swamp environments and
associated flora on a geological time scale. Such
changes are reflected in the pronounced variations
observed in diterpane fingerprints over depth intervals on the order of 100 m or less in Rakopi coals
from Tane-1. Hence, generated oils can be expected to reflect a mixture of these biomarker variations, given the vertical extent of source rock units
likely to be in the oil window at a given time.
Correlation of oils with their potential source
rocks is most readily achieved by following higherplant evolution. Although angiosperms first
appeared in the Cretaceous and rapidly rose to
dominance on a global basis, in New Zealand the
pollen record shows that gymnosperms, particularly podocarps and araucarians, dominated the higher-plant communities of coastal flood plains during
the Late Cretaceous and were still major members
throughout the Paleocene, but had become subordinate to angiosperms by the Eocene (Mildenhall,
1980). Biomarker distributions ref lect these
changes in the flora. Coals and interbedded shales
from the Late Cretaceous Rakopi Formation exhibit
very high levels of gymnosperm-derived diterpanes, often dominated by isopimarane, and
extremely low levels of 18(H)-oleanane [2% relative to 17(H)-hopane from m/z 191 response]. In
contrast, Eocene Mangahewa and Kaimiro formation coals exhibit relatively high levels of
angiosperm-derived triterpenoids, particularly
oleanoids, ursanoids, and lupeoids, and muchreduced levels of diterpanes. Paleocene coals in
Kapuni Deep-1 (i.e., belonging to Kapuni Group
cycle A) appear to exhibit biomarker source characteristics more like those of Late Cretaceous coals
(Johnston et al., 1990).
The biomarker characteristics discussed above can
be represented by an angiosperm/gymnosperm index
(AGI), which is calculated from the relative amounts
of selected diterpanes and triterpanes in m/z 123 and
191 mass chromatograms, respectively. A plot of log
Figure 6Differentiation of main oil families from AGI vs. log TMI for various sediment samples, labeled
biomarker parameters derived from m/z 191 mass chro- according to formation, and oils is presented in Figure
matograms (see Table 3 for description of parameters).
7 (see Materials and Methods section for evaluation of
indices). As expected, coals and shales of the Mangahewa and Kaimiro formations generally plot at high
Canterbury (Gibbons and Fry, 1986), and Great AGI, whereas those of the Rakopi Formation plot at
South (Raine et al., 1993). This shale is present in the low AGI. The lowest AGI values (<0.5) for Eocene sedAriki-1 well, and representative TOC and Rock-Eval iments were recorded for samples from the Kaimiro
data from other locations are presented in Table 5 in and deepest Mangahewa formations in Waihapa-1,
the absence of data from Ariki-1.
and are consistent with the expected slightly lower
Although there are no perfect matches between angiosperm contribution in the early Eocene comany particular source rock sample and any oil, there pared with the mid-to-late Eocene.

Killops et al.

1573

Table 5. Mean Total Organic Carbon and Rock-Eval Data for Formations in Various Regions of the
Taranaki Basin*
Field/Region

Formation

TOC
(%)

S2
()

HI
()

Kaimiro

Mangahewa
Turi + Kaimiro
Mangahewa
Farewell
Mangahewa (upper)
Mangahewa + Kaimiro
Farewell
North Cape
Rakopi
Turi
North Cape
Rakopi
Turi
Turi (Waipawa)
Turi (Waipawa)

13.9
1.6
13.0
1.7
9.7
10.4
9.4
0.9
8.9
0.6
0.4
4.9
1.9
3.2
6.1

42.3
2.3
27.5
1.5
29.8
21.8
16.1
0.8
23.4
0.1
0.1
4.9
3.2
7.5
19.9

239
120
137
85
230
194
174
75
243
16
25
118
138
236
323

Kapuni
Kupe
McKee
Maui-4

Tane-1
Witiora-1
Angora Stream
Galleon-1

*Waipawa Black Shale samples are from the East Coast Basin (Angora Stream), after Leckie, personal communication; and Canterbury Basin (Galleon-1
well), after Gibbons and Fry, 1986; other data after Analabs, 1984.

For Rakopi and Mangahewa/Kaimiro formation


sediments, TMI values are generally greater than 2.
Within predominantly terrestrial sediments, however, varying marine influences are suggested by the
presence of marine organisms (e.g., dinoflagellates)
and general stratigraphy (Flores et al., 1993), and
are reflected in biomarker distributions. For example, a bulk upper Rakopi sample from Fresne-1
exhibits a TMI of 2.4, whereas coal isolated from
this sample has a TMI of 8.4. It appears that the
shales interbedded with coals in such sediments
represent relatively brief marine incursions during
a phase that is basically regressive but in which the
input of terrestrial organic detritus is still dominant. Rakopi samples from Maui-4 have lower TMI
values (<6) than those from Tane-1, suggesting a
greater marine inf luence in the Maui-4 region.
These inferences of marine influence in Maui-4 are
consistent with the sulfur isotopic distributions discussed in the previous section (Table 4).
AGI values for Turi Formation samples range
from intermediate to high (ca. >0.1), as expected
for Paleogene samples, whereas TMI values are low
(ca. <2), which is consistent with significant
marine contributions. Farewell Formation samples
exhibit intermediate to low AGI values, again as
expected; the lowest value, similar to those of the
youngest Rakopi Formation coals, is recorded for a
sample from near the base of the main Farewell
coal unit in Kapuni Deep-1 (within the D cycle,
4693 m). The deepest Farewell samples in Kapuni
Deep-1 (5308 and 5520 m) show an increasingly
marine character, with TMI values of 1.0 or less and
high 23T/dH values (>3.5).

A strong correlation is apparent between oils of


the northern part of the Tarata thrust zone (McKee3A, Pouri-1A, Toetoe-1, and Urenui-1, Figure 7) and
Eocene coals. There also appears to be a major
Eocene contribution to oils from Ahuroa, Moturoa,
Kaimiro, Ngatoro, Stratford, and Tariki fields and oil
from the Tikorangi reservoir in the Waihapa field
(Ahuroa-1, Kaimiro-1, Moturoa-2, Stratford-1, Tariki1A, and Waihapa-2, Figure 7). However, fairly low
TMI values for the oils from Moturoa, Kaimiro, Ngatoro, and Stratford fields (Table 3) suggest a greater
marine inf luence than for the other Eocenesourced oils. The oils from Mangahewa-1 and the
Kaimiro reservoir in Waihapa-1A appear to have a
greater Paleocene contribution than other oils from
just west of the Tarata thrust zone (Figure 7). Maui
oils correlate well with Rakopi coals, whereas the
oil from Moki-1 may contain Late Cretaceous and
Paleocene contributions (Figure 7). Oils of the
Kapuni and Kupe South fields plot at intermediate
AGI, suggesting Paleocene or mixed Eocene/Late
Cretaceous sources, predominantly coals from the
high TMI values (Figure 7).
Levels of trans,trans,trans-bicadinane relative to
hopane were assessed from m/z 369 and 412
responses for some representative oils (Table 6),
and are consistent with AGI values. Oils with significant contributions from Eocene coals (e.g., in
McKee and Kaimiro fields) exhibit the highest relative levels, whereas those of predominantly Paleocene origin (e.g., Kupe South field) contain
markedly lower levels and Cretaceous-sourced oils
(e.g., Maui field) contain the lowest amounts (at
the limit of detection). Bicadinane levels do not

1574

Oil Generation in the Taranaki Basin

(Castlepoint; Table 4) are similar and consistent


with a depositional environment near the shelf
break or on the slope, probably under a highly productive photic zone. The development of anoxicity
in surface sediments is indicated by an intense -log
signal for the shale in both Ariki-1 and Galleon-1
(Canterbury Basin). Although the shale is only
some 10 m thick in Ariki-1 and 7 m thick in
Galleon-1, it has a high petroleum genetic potential
(Table 5). A contribution from an equivalent of this
shale is also apparent in the oil show from
Pukearuhe-1. It should be noted that none of the
sediment samples analyzed in this study exhibits an
exclusively marine signature, owing to the input of
varying quantities of terrestrial detritus to shelf
environments.
OIL GENERATION AND EXPULSION FROM
NEW ZEALAND COALS

Figure 7Logarithmic plot of angiosperm/gymnosperm


vs. terrestrial/marine indices for oils and sediment
extracts (labeled by formation: F = Farewell, K =
Kaimiro, M = Mangahewa, N = North Cape, R = Rakopi, T
= Turi). Circled areas represent relative bacterial contributions to oils (hopane:sterane ratio). (See Table 2 for
well abbreviations.)

parallel those of oleanane exactly: for example, the


BC:O ratio for Maui-4 oil is very low compared with
that for McKee-3A (Table 6). Similarly, although
oleanane is present in a mature Rakopi coal sample
from Fresne-1 (4% relative to 30 from m/z 191
response), no bicadinane was detected, suggesting
no more than minor occurrence of dipterocarptype flora in the Late Cretaceous in New Zealand.
Kora-1 and Tangaroa-1 oils exhibit good biomarker correlation with a late Paleocene shale (Waipawa
Black Shale equivalent) at 41204130 m in Ariki-1
(Figure 5), although the shale is less mature than
the oils. The marine biomarker signature is augmented by varying levels of angiosperm-derived
biomarkers [e.g., wax n-alkanes and 18(H)oleanane]. Sulfur isotopic ratios for Kora-1 oil
(Table 2) and the shale in the East Coast Basin

The Taranaki and Gippsland (southwest Australia) basins were relatively close together at the
end of the Cretaceous and so, not surprisingly, their
terrestrial oils have compositional affinities, reflecting general similarities in depositional environments associated with lower coastal plain swamp
ecosystems (Thomas, 1982). These include the
presence of oleanane (O), C29 neohopane (29N),
diahopane (30D), and gymnosperm-derived diterpanes; abundant diasteranes; limited amounts of tricyclic terpanes; high hopane:sterane ratios, wax
contents, and pristane:phytane ratios (4); and low
sulfur contents (Philp and Gilbert, 1986; Alexander
et al., 1987). Low sulfur levels are consistent with
deposition in freshwater to low-salinity environmentstypical conditions for the source rocks of
high-wax oils, most of which appear to have been
deposited in the Cretaceous and Tertiary (Hedberg,
1968; Gould, 1980). It is now recognized that oilprone coals were deposited under conditions in
which microbial degradation of woody tissue could
occur, resulting in concentration of the more resistant, hydrogen-rich components such as cuticles
and resin bodies (Kirkland et al., 1987; Powell,
1988; Powell et al., 1991). Such conditions appear
to be mostly associated with temperate rather than
tropical climates (Thomas, 1982), as during source
rock deposition in the Taranaki Basin (Mildenhall,
1980).
The oil potential of the hydrogen-rich, Middle
Jurassic Walloon and Tertiary Latrobe group coals
of Australia is attributed to high exinite content, in
the form of suberinite, cutinite, and, in particular,
resinite (Thomas, 1982; Khorasani, 1987). However, New Zealand coals are vitrinite-rich and contain
little exinite (Newman and Newman, 1982). The
vitrinite is mainly in the form of the amorphous,

Killops et al.

Table 6. Distribution of trans,trans,transBicadinane (BC) in Some Taranaki Oils Relative to


17(H)-Hopane (30) and 18(H)-Oleanane
Based on m/z 369 Responses
Oil

BC
30

BC
O

Moturoa-2
Kaimiro-1
Kaimiro-2
McKee-3A
Waihapa-1A
Waihapa-2
Kupe South-1
Maui-1
Maui-4

2.99
2.28
1.36
1.07
nd*
0.75
0.29
nd
0.13

7.84
5.18
3.55
1.64
nd
2.26
0.92
nd
0.29

*nd = BC not detected.

hydrogen-rich, submaceral desmocollinite, which is


probably primarily derived from lipid-rich, leafcuticular membranes (comprising epicuticular wax
and the cuticle proper; Holloway, 1982) and bacterial cell walls and membranes. These components
are known to be resistant to bacterial degradation
during diagenesis (e.g., Nip et al., 1986; Tegelaar et
al., 1989; Le Berre et al., 1991), and can be major
components in lignites [e.g., Yallourn lignite
(Noble et al., 1985) and some Northland lignites
(Mildenhall, 1988)]. It is the epidermal cell wall
architecture that confers the characteristic shape of
cutinite, but once the cell wall has been degraded
it is possible for the cuticular material to adopt
whatever shape is dictated by external pressures
and become unrecognizable (i.e., amorphous). A
similar fate may also be possible for suberinite, and
bacterial cell walls and membranes (containing
large amounts of hopanoids) would also be expected to yield amorphous material. Hence, not all oilprone coals contain large amounts of suberinite
and cutinite (Powell et al., 1991), and hydrocarbon
potential, when not associated with high liptinite
content, correlates with hydrogen-rich vitrinite
abundance (Bertrand, 1989). The presence of leafcuticular membrane material would account for
the waxy nature of Taranaki oils and their abundant
higher plant diterpanes and triterpanes (Aplin et
al., 1963; Cambie and Weston, 1968; Killops and
Frewin, 1994). In addition, the highly resistant
cuticular component cutan is probably responsible
for the significant polymethylene signal observed
in 13C NMR studies of Late Cretaceous and Tertiary
New Zealand coals (Collen et al., 1988).
It has been proposed that, although oil can be
generated by New Zealand coals at a rank of ca.
0.7% Ro, it is not expelled until significantly higher
maturity levels of ca. 1.0% Ro are reached at burial
depths in excess of 5 km, because of the

1575

chemisorption and molecular sieve properties of


coals (Cook, 1987; Cook, 1988; Johnston et al.,
1991). Because exploration wells have not penetrated source rock horizons exhibiting biomarker
maturity levels as high as those recorded for oils, it
has not proven possible to establish exact generation and expulsion maturity thresholds. The most
useful molecular maturity parameter from the saturated hydrocarbons over the maturity range in
question, the relative amount of 5(H),14(H),
17(H)-24-ethylcholestanes compared with 5(H),
14(H),17(H)-24-ethylcholestanes, can vary in
value at the end of diagenesis as a result of sourcerelated effects (Peakman et al., 1989), and so it is
not possible to obtain absolute maturity estimates
of the expulsion threshold from this ratio in oils
alone. However, such estimates are possible when
potential source rock data are available, permitting
discrimination between source- and maturity-related effects. Given that McKee field and related oils
are largely sourced by Eocene coals, that Kapuni/
Kupe oils are Paleocene-sourced, and that Maui oils
are predominantly sourced by Late Cretaceous
coals, an approximation of expulsion thresholds
can be obtained.
In Figure 8, the extent of isomerism at C-14 and
C-17 in 5(H)-24-ethylcholestanes [/( + )] is
plotted against depth for Mangahewa Formation
sediments from McKee-1 and the maturity trend is
extrapolated to the mean value exhibited by oils
from the area, which is an approximation of the
expulsion threshold. The corresponding vitrinite
reflectance and rank (S) values for the source rock
at oil expulsion are obtained by reference to the
plots for McKee-1 coals, based on established depth
trends (Lowery, 1988; Sykes et al., 1992). The
expulsion threshold is ca. 0.8 0.1% Ro, or ca. 13.2
0.5 R(S), at a depth of just over 4000 m in McKee1 [allowing for uncertainties in vitrinite reflectance
and rank (S) trends].
A similar biomarker maturity approach for
Kapuni/Kupe oils and Farewell Formation sediments in Kapuni Deep-1 yields expulsion-threshold
values of 0.85 0.1% Ro and 14.6 1.0 R(S) at ca.
5500 m depth (Figure 8). For Maui oils, as will be
seen, the exact source area is unknown, but expulsion characteristics can be examined for Rakopi
Formation sediments in Maui-4 and Tane-1 if it is
assumed that Rakopi coals throughout the basin
have similar chemical characteristics. The oil-expulsion threshold appears to be ca. 0.9 0.1% Ro or
13.8 0.5 R(S) in Tane-1 and 0.9 0.1% Ro or 13.3
0.5 R(S) in Maui4 (Figure 8). These values are
self-consistent and do not appear to be affected by
differing thermal regimes for the two regions and
the loss of ca. 1300 m of sediment in the Maui-4
area during late Miocene uplift and erosion.
Although the Maui oils appear slightly more mature

1576

Oil Generation in the Taranaki Basin

Figure 8Depth correlations of the extent of isomerism at C-14 and C-17 in 5(H)-24-ethylsteranes with vitrinite
reflectance (Ro) and rank (S) [R(S)] for Mangahewa Formation sediments from McKee-1 (and Urenui-1), Farewell
Formation sediments from Kapuni Deep-1, Rakopi Formation sediments from Tane-1, and Rakopi Formation sediments from Maui-4. [Ro and R(S) depth calibration after Sykes et al., 1992; Ro data after Lowery, 1988.]

than the northern Tarata thrust zone oils on the


basis of implied vitrinite reflectance values for their
source rocks during expulsion, rank (S) values are
similar. Rank (S) (also known as Suggate rank) more
accurately represents absolute maturity (e.g., varying heating rates) and does not exhibit the sourcerelated effects of Ro measurements (e.g., Suggate
and Boudou, 1993).
It appears that the oil-expulsion threshold for
New Zealand coals occurs at ca. 12.514 R(S),
although it must be remembered that this value
represents an average rank for all the source rock

strata that have contributed to the pooled oil and


so is likely, if anything, to overestimate the maturity
at which expulsion commences. However, it is consistent with peak hydrogen index (HI) for New
Zealand coals being reached around 13 R(S) and
thereafter decreasing, as a consequence of the
expulsion of generated hydrocarbons (Bertrand,
1989; Sykes et al., 1992). There is no evidence that
biodegradation has occurred in oil accumulations
to a degree that would affect the /( + ) sterane ratio (which would, in any event, tend to cause
an overestimation of the expulsion threshold;

Killops et al.

1577

Figure 9van Krevelen-type diagram for the New Zealand coal band showing Suggate isorank contours (elemental
compositions expressed on nitrogen-free, sulfur-free, and dry-mineral-matter-free basis; after Suggate, 1959; Suggate, personal communication) and generalized representation of fluid generation from a typical New Zealand coal
with increasing rank (S) per tonne of carbon in lignite of initial R(S) ca. 4. [Fluid volumes based on typical Mangahewa reservoir densities in Kapuni field of 0.6 gcm3 for CO2 and 0.8 gcm3 for oil, represented by (CH2)n.]

McKirdy et al., 1983). The bulk chemical constitution and petroleum potential of Pakawau and
Kapuni group coals are similar and fall within the
maturity trend of the New Zealand coal band on
van Krevelen-type diagrams (Figure 9), and so similar petroleum generation and expulsion characteristics are to be expected. Typically, the hydrogen
index (HI) of New Zealand coals rises to ca. 300 at
rank (S) 12.514 (vitrinite ref lectance of ca.
0.70.9% Ro), and the Rock-Eval S2 parameter is ca.
200 (Suggate and Boudou, 1993).
It is possible to model fluid evolution from a typical New Zealand coal with increasing rank (S), as
shown in Figure 9, based on an established kinetic
model for type III kerogen (Burnham and Sweeney,
1989; Sweeney, 1990; Sweeney and Burnham,
1990). Oil expulsion appears to occur when ca.
30% of the genetic potential has been realized,
which occupies ca. 10% of the source rock volume
at typical subsurface conditions. Applying this 30%
threshold to more sophisticated thermal history
and kinetic modeling (Armstrong et al., 1994)
yields expulsion-threshold depths very close to
those obtained above from the biomarker maturity
approach.
The potential for early generation of oil from
hydrogen-rich coals, such as those of the Taranaki
Basin, and the limited extent of absorption and

adsorption processes that might hinder expulsion,


have recently been recognized (e.g., Noble et al.,
1991; Sandvik et al., 1992). Clearly, primary migration may occur at significantly shallower depths
than has been generally thought possible in the
Taranaki Basin. This conclusion applies to the average coal on the New Zealand coal band, and there
will be some variation in oil generation and expulsion characteristics depending on the plant and tissue types preserved in a given peat, which are related to depositional environment, as found for the
coals of the Talang Akar Formation of Java (Noble
et al., 1991).
Large quantities of carbon dioxide are associated
with gas/condensate accumulations within the
upper part of the Mangahewa Formation in
onshore wells where this formation has not yet
entered the oil window. For example, ca. 40% of
the gas is CO2 in the Kapuni field (McBeath, 1977).
Compositional and isotopic studies suggest that the
CO 2 originates from carboxyl group elimination
from coals during the lignite to early high-volatile
bituminous coal stages (Boudou et al., 1984;
Giggenbach et al., 1993). Various removal processes and dilution with methane from more deeply
buried coals probably lead to a quite rapid decline
in CO2 concentrations once CO2 generation ceases,
but large amounts are still being evolved by the

1578

Oil Generation in the Taranaki Basin

Figure 10Cross sections of onshore and near-shore areas of the Taranaki Basin (after King et al., 1991), showing
the main petroleum source rock formations (F = Farewell, K = Kaimiro, M = Mangahewa, N = North Cape, R =
Rakopi, T = Turi) and their spatial relationships to oil accumulations (black lenses). (See Figure 1 for location of
cross sections. Medium tone = organic-rich units, partial tone = reduced organic richness, dark tone = units that
have entered oil-expulsion window. Random ticks indicate seismic basement.)

onset of significant oil generation (Figure 9). Under


typical subsurface temperature and pressure
regimes for accumulations in the Taranaki Basin,
CO 2 (and mixtures with CH 4 ) is a supercritical
fluid, possessing considerable solvating potential
for hydrocarbons, which aids the primary migration of oil (see also McKirdy and Chivas, 1992). The
large volume of CO2 evolved prior to oil generation
would also be anticipated to aid primary migration
indirectly, by effecting microfracturing of the
source rock. It may also be responsible for remobilizing oils, generated at greater depth, that have

become trapped beneath the upper Mangahewa


coal unit. The f luid is less dense than oil and,
together with methane evolved from deeper
Rakopi coals, may displace oil past reservoir spill
points, causing what may be termed tertiary migration into shallower traps. This process may account
for the dominance of gas (under surface conditions) in upper Mangahewa reservoirs of the
Kapuni and Kaimiro fields, and the accumulation of
oils/condensates in late Miocene Mount Messenger
Formation sandstones in the Kaimiro and Ngatoro
fields (Table 1 and Figure 2).

Killops et al.

1579

Figure 11Cross sections of offshore regions of the Taranaki Basin. (See Figure 1 for locations of cross sections and
Figure 10 legend for key to symbols. Dots immediately above seismic basement in Tangaroa-1 represent oil show.)

BASIN-WIDE OIL GENERATION AND


EXPULSION CHARACTERISTICS
The cross sections in Figures 10 and 11 show the
main source rock units in various regions of the
basin and horizons likely to have entered the oil
window at some time (i.e., generation and expulsion), based on biomarker maturity parameters that
are in agreement with preliminary kinetic models
(Armstrong et al., 1994). The composition of oil
accumulations generally reflects the largest volume
of source rock in the vicinity that has entered the
oil window within the last 10 Ma (cf. Armstrong et
al., 1994). It appears that oils of distinctly Rakopi

origin, with the exception of those from the Maui


field, Maui-4, and possibly Moki-1, have largely
escaped from the system, being mostly expelled
more than 20 Ma, prior to the development of
traps. The expulsion window occurs at a presentday depth of 5.05.5 km in the low-heat-flow area
in the southeast region of the basin (including
Kapuni and Kupe South fields; cross sections AA
and BB, Figure 10), at ca. 4 km in the higher-heatf low area of the northern Taranaki Peninsula
(encompassing Moturoa, Kaimiro, and McKee
fields; cross section CC, Figure 10) and at intermediate depths over much of offshore Taranaki Basin
(depending on heat flow, Figure 1).

1580

Oil Generation in the Taranaki Basin

Eocene coals have not reached sufficient maturity to contribute to the oils of the Kapuni and Kupe
South fields (cross section AA, Figure 10), and so a
mixed Eocene/Late Cretaceous origin seems far less
likely than a predominantly Paleocene (i.e., Farewell
Fm.) coal source for the Kapuni and Kupe South
oils. Rakopi coals probably make a significant contribution to hydrocarbon gas accumulations in the
Kapuni field (cf. Armstrong et al., 1994), whereas
immature Mangahewa coals have generated the
existing CO 2. The Farewell Formation becomes
quite organic poor to the south of the Kapuni field,
other than the coals of the Puponga Member, which
spans the base of the Farewell Formation and the
top of the North Cape Formation (cross section BB,
Figure 10). It would appear that the Kupe South
field oils are sourced mainly by the Farewell Formation in the Kapuni field region. However, the
Kapuni/Kupe oils exhibit a lower marine contribution than the sediments analyzed from depths
greater than 5000 m in Kapuni Deep-1, which suggests that if this interval has sourced the oils, coal
bands are present in the drainage area and have a
significantly greater oil potential than the marineinfluenced horizons sampled. The Rakopi Formation must lie within the oil window in the region of
Toru-1 (cross section BB, Figure 10), but there is no
indication that expelled oil has been trapped. Here,
as elsewhere in the basin where exploration wells
have not penetrated the Rakopi Formation, the presence of coals can be inferred only from seismic
reflection characteristics.
In the Waihapa field there are clearly compositional differences between the oils of the Kaimiro
and Tikorangi formation reservoirs. The Kaimiro
Formation oil appears to be predominantly Paleocene sourced, like the Kapuni/Kupe oils. However, biomarker maturity parameters suggest that the
oil-expulsion threshold is a little shallower (at ca.
5000 m) than in the Kapuni field, which is consistent with a slightly higher heat flow (Figure 1), and
that Mangahewa Formation coals near the basement-wall thrust have entered the expulsion window (cross section AA, Figure 10). Mangahewasourced oil could not accumulate in the Kaimiro
Formation reservoir, but would be anticipated to
contribute to the Tikorangi reservoir, accounting
for the observed compositional differences
between oils from the two reservoirs.
The oils of the northern Tarata thrust zone, such
as those from the McKee field, appear to be
sourced mainly from Mangahewa coals at the base
of the overthrust, the oils having migrated up-structure to reservoirs in the McKee Formation sandstones (cross section CC, Figure 10). There is no
reason to suspect that oleanane and other
angiosperm-derived biomarkers have been
entrained by oil generated from deeper sources

during its vertical migration through Eocene coals


(Philp and Gilbert, 1986; Johnston et al., 1991),
because diterpane levels relative to hopanes are too
low for a Rakopi source and kinetic modeling indicates that Rakopi coals expelled oil in the Paleocene and are now gas prone (Armstrong et al.,
1994).
To the west of the Tarata thrust zone, a gradual
change in oil composition can be observed on moving northwest through Stratford, Ngatoro, Kaimiro,
and Moturoa fields. An increasing marine contribution is observed, which is expected from the
increasingly marine character of Paleogene (Turi
Fm.) units (cross section CC, Figure 10). Although
not particularly organic-rich, these units are quite
thick. However, the dominant inf luence on oil
composition is from adjacent, mature, Eocene coals
(mainly Mangahewa, because the Kaimiro Formation is relatively organic lean in this region),
although it is probable that mature, terrestrial, Paleocene (Farewell Fm.) strata immediately to the
west of the Tarata thrust zone also contribute. Both
Mangahewa and Mount Messenger formation reservoirs in the Kaimiro field lie above the source rock
strata (cross section CC, Figure 10), and so sourcerelated compositional differences of the type
described above for the two Waihapa field reservoirs would not be expected. The oils from the two
Kaimiro field reservoirs are compositionally alike
and so could have a common source. Gas, particularly CO 2 released from Mangahewa Formation
coals, may be responsible for displacing oil from
the Mangahewa reservoir into the shallower Mount
Messenger reservoir in the Kaimiro field.
Only thin and immature Rakopi sediments are
encountered in the Maui wells and Moki-1 (cross
section AA, Figure 10). Preliminary thermal modeling indicates that, in areas where Late Cretaceous
sediments are thickest, the Rakopi Formation first
entered the oil window in the Paleocene and is
now gas prone (Armstrong et al., 1994), and so
expelled oil probably escaped. The source area for
the Maui oils is most likely to be found where the
Rakopi Formation has entered the oil window
within the last 10 Ma. Two possible source
regions that satisfy this criterion have previously
been proposed, one lying to the east, across the
Cape Egmont fault (Thrasher, 1990), and the other
lying to the northeast in the vicinity of New Plymouth (Haskell, 1991, 1992). It is not possible to
eliminate either of these on the basis of data
presently available. Anomalously high 3 He/ 4 He
ratios recorded in gases from both Maui and New
Plymouth areas do not necessarily establish a link
(Giggenbach et al., 1993). This mantle-gas signature could be the result of crustal fractures in the
Quaternary rift zone identified under the Central
Graben and the southern part of the North

Killops et al.

Graben (King and Thrasher, 1992), and so may be


a feature of the entire rift zone (Allis, personal
communication).
Rakopi coals below a depth of 4 km may have
entered the oil window in the vicinity of Maui-4
(cross section DD, Figure 11), but were removed
from it again by ca. 1300 m of post-Miocene uplift.
Deeper into the Southern Inversion Zone, where
up to 3 km of uplift and erosion has occurred in the
Quaternary (Allis et al., personal communication;
Kamp and Green, 1990), source rocks that were
previously in the oil window had their maturation
reactions suspended by the uplift (e.g., Rakopi
coals in Fresne-1; cross section EE, Figure 11). In
contrast, on the Western Stable platform it appears
that base-Rakopi coals may only just be approaching the expulsion window (e.g., Tane-1; cross section FF, Figure 11).
The probable source of the Kora-1 oil, the
Waipawa Black Shale equivalent, probably enters
the oil window at 5.05.5 km burial depth to the
east of Kora-1 well (cross section GG, Figure 11),
and migration up into the Kora volcanic structure
appears straightforward (cross section HH, Figure
11). From north-south and east-west stratigraphic
cross sections for the region (King et al., 1991), the
likely drainage area for the Kora oil is ca. 500 km2.
Assuming that the shale is consistently 10 m thick,
a gas:oil ratio of 0.3 as for the shale from Galleon-1
(Gibbons and Fry, 1986), a porosity of 10% and wet
density of 250 kgm3 (typical for the burial depth),
and using the Rock-Eval data in Table 5, the oil
potential of this source rock unit is ca. 155 Mt or
ca. 170 106 m3 under surface conditions (ca. 200
106 m3 under reservoir conditions of density ca.
740 kgm3 at 17.8 MPa and 68C). The volume of
oil in the Kora-1 reservoir is estimated at 5 106 bbl
(McManamon, 1993), or just under 10 6 m 3. This
represents no more than ca. 1% of the genetic
potential of the assumed drainage area, and so the
shale is a reasonable source for the Kora oil on
quantitative grounds. It appears that the Waipawa
Black Shale equivalent extends northward into the
Northland Basin and is a potential oil source rock
there (Isaac and Herzer, 1994).
CONCLUSIONS
It is possible to determine the relative terrestrial
and marine contributions to source rocks in the
Taranaki Basin from the carbon number distributions of steranes (terrestrial/marine index). In addition, the trend in increasing importance of
angiosperms in coastal flood plain plant communities from the Late Cretaceous to the Eocene can be
followed by monitoring the relative proportions of
angiosperm-derived triterpanes and gymnosperm-

1581

derived diterpanes (angiosperm/gymnosperm


index), yielding age-related source rock information. The use of a plot of angiosperm/gymnosperm
index vs. terrestrial/marine index permits the
sources of Taranaki Basin oils to be determined
with reasonable accuracy, especially when combined with maturity considerations based on heatflow data and derived kinetic models of petroleum
generation. Coals of the Pakawau and Kapuni
groups are the main oil-source units, but marineinf luenced shales also contribute to varying
extents.
Oil composition ref lects the distribution of
mature source rocks near reservoirs, given that
most Rakopi-sourced oil may have escaped prior to
Neogene trap formation. Distinctly Rakopi-sourced
oil is represented only by the Maui accumulations
and contributions to the Moki field oil. There is the
possibility, however, that some could be trapped in
deeper reservoirs where secondary porosity has
developed. Kapuni and Kupe South oils seem to be
predominantly sourced by Paleocene (Farewell
Fm.) sediments. There is also a Paleocene contribution to oils in the Waihapa field, at the south end of
the Tarata thrust zone, although the Neogene reservoir in this field has also received a contribution
from Eocene coals. In the north of the Taranaki
Peninsula there is a high-heat-flow area centered on
New Plymouth, and mature source rock units lie
nearer the surface.
Oils of the northern part of the Tarata thrust
zone originate mainly from Eocene coals of the
Mangahewa Formation and, to a lesser extent, the
Kaimiro Formation. Toward New Plymouth, coalrich strata within the Kaimiro and Farewell formations grade into relatively organic-lean, marine
shales of the Turi Formation, which appear to contribute to the oils of the area (e.g., Kaimiro and
Moturoa fields). Mangahewa coals are still present
but are mostly early mature and so their contributions are relatively limited compared with those of
the northern Tarata thrust zone. There is probably
also a significant contribution from mature
Farewell coals lying immediately to the east of the
New Plymouth/Kaimiro region. In the northwest of
the basin an equivalent of the late Paleocene
Waipawa Black Shale is present and has been
buried sufficiently deeply to source some oil, such
as that in the Kora volcanic structure.
Hydrogen-rich, Late Cretaceous and Tertiary
New Zealand coals are capable of generating and
expelling oil by rank (S) 12.514.0 throughout the
Taranaki Basin, based on biomarker maturity studies and the implications of compositional changes
occurring through the New Zealand coal band. The
corresponding vitrinite reflectance range is ca.
0.70.9% R o , over a depth range of 4.05.5 km,
depending on thermal regime. The oil potential of

1582

Oil Generation in the Taranaki Basin

New Zealand coals (and associated shales) is


attributable to the submaceral desmocollinite. This
amorphous material appears to be composed mainly of leaf-cuticular membranes (accounting for high
saturates and wax content and the presence of
abundant higher-plant diterpanes and triterpanes)
and bacterial remains (also aliphatic-rich and contributing the hopane content). Such material is
characteristically deposited under relatively oxidizing conditions in temperate swamp environments
in which microbial degradation of the bulk of lignified, cellulosic tissue occurs, concentrating the
more resistant, aliphatic-rich components. Primary
migration of oil from coal appears to be aided by
the generation of large volumes of supercritical carbon dioxide from coals, creating expulsion pathways by microfracturing source units and transporting the first phase of generated oil in solution.

REFERENCES CITED
Alexander, R., R. A. Noble, and R. I. Kagi, 1987, Fossil resin
biomarkers and their application in oil to source-rock correlation, Gippsland Basin, Australia: APEA Journal, v. 27, p. 6372.
Analabs, 1984, Petroleum geochemistry of the Taranaki Basin:
IGNS Petroleum Report 1013, Institute of Geological and
Nuclear Sciences, Lower Hutt.
Analabs, 1989, Petroleum Geochemistry: Okoki-1: IGNS Petroleum
Report 1495, Institute of Geological and Nuclear Sciences,
Lower Hutt.
Analabs, 1991, Petroleum Geochemistry: Toru-1: IGNS Petroleum
Report 1668, Institute of Geological and Nuclear Sciences,
Lower Hutt.
Aplin, R. T., R. C. Cambie, and P. S. Rutledge, 1963, The taxonomic distribution of some diterpene hydrocarbons: Phytochemistry, v. 2, p. 205214.
Armstrong, P. A., D. S. Chapman, R. H. Funnell, R. G. Allis, and P. J.
J. Kamp, 1994, Thermal state, thermal modelling and hydrocarbon generation in the Taranaki Basin: Proceedings of New
Zealand Petroleum Conference 1994, Ministry of Commerce,
Wellington, p. 289-307.
Baker E. A., 1982, Chemistry and morphology of plant cuticular
waxes, in D. F. Cutler, K. L. Allen, and C. E. Price, eds., The
plant cuticle: Linnean Society, London, Academic Press,
p. 139165.
Bennett, C., R. Gregg, and P. King, 1992, Petroleum geology of the
Taranaki Basin, with emphasis on the north-eastern quadrant:
Petroleum Exploration New Zealand News, May, p. 1425.
Bertrand, P. R., 1989, Microfacies and petroleum properties of
coals as revealed by a study of North Sea Jurassic coals: International Journal of Coal Geology, v. 13, p. 575595.
Boudou, J.-P., B. Durand, and J.-L. Oudin, 1984, Diagenetic trends
of a Tertiary low-rank coal series: Geochemica et Cosmochimica Acta, v. 48, p. 20052010.
Brassell, S. C., G. Eglinton, and F. Jiamo, 1986, Biological marker
compounds as indicators of the depositional history of the
Maoming oil shale: Organic Geochemistry, v. 10, p. 927941.
Burnham, A. K., and J. J. Sweeney, 1989, A chemical kinetic model
of vitrinite maturation and reflectance: Geochimica et CosmochimicaActa, v. 53, p. 26492657.
Burwood, R., P. Leplat, B. Mycke, and J. Paulet, 1992, Rifted margin source rock deposition: a carbon isotope and biomarker
study of a West African Lower Cretaceous lacustrine section:
Organic Geochemistry, v. 19, p. 4152.
Cambie, R. C., and R. J. Weston, 1968, Chemotaxonomy of the

New Zealand Podocarpaceae: Journal of New Zealand Institute


of Chemistry, v. 32, p. 105121.
Collen, J. D., J. H. Johnston, P. J. Dunn, and R. H. Newman, 1988,
13C NMR spectra from Upper Cretaceous and lower Tertiary
coal measures, Taranaki Basin, New Zealand: Geology Board of
Studies Publication, no. 2, Victoria University, Wellington.
Collier R. J., and J. H. Johnston, 1991, The identification of possible source rocks, using biomarker geochemistry, in the Taranaki Basin, New Zealand: Journal of Southeast Asian Earth Sciences, v. 5, p. 231239.
Cook, R. A., 1987, The geology and geochemistry of crude oils and
source rocks of western New Zealand: New Zealand Geological
Survey Petroleum Report 1250, Institute of Geological and
Nuclear Sciences, Lower Hutt.
Cook, R. A., 1988, Interpretation of the geochemistry of oils of
Taranaki and west coast region, western New Zealand: Energy
Exploration and Exploitation, v. 6, p. 201212.
Czochanska Z., T. D. Gilbert, R. P. Philp, C. M. Sheppard, R. J.
Weston, T. A. Wood, and A. D. Woolhouse, 1988, Geochemical
application of sterane and triterpane biomarkers to a description of oils from the Taranaki Basin in New Zealand: Organic
Geochemistry, v. 12, p. 123135.
De Grande, S. M. B., F. R. Aquino Neto, and M. R. Mello, 1993,
Extended tricyclic terpanes in sediments and petroleums:
Organic Geochemistry, v. 20, p. 10391047.
Flores, R. M., J. M. Beggs, and P. R. King, 1993, Sedimentology of
tide-dominated reservoir sandstones in the Eocene Kapuni
Group, Taranaki Basin, New Zealand (abs.): AAPG 1993 Annual
Convention, p. 102.
Frankenberger A., R. R. Brooks, H. Varela-Alvarez, J. D. Collen,
R. H. Filby, and S. L. Fitzgerald, 1994, Classification of some
New Zealand crude oils and condensates by means of their
trace element contents: Applied Geochemistry, v. 9, p. 6571.
Gibbons M. J., and S. Fry, 1986, A geochemical study of the
Galleon-1 well, Canterbury Basin, offshore New Zealand: IGNS
Petroleum Report 1146, Institute of Geological and Nuclear Sciences, Lower Hutt.
Gibbons M. J., H. A. Bockmeulen, and S. OReilly, 1981, Petroleum
geochemistry of the Shell BP Todd well: Tangaroa-1, offshore
Taranaki, New Zealand: IGNS Petroleum Report 793, Institute
of Geological and Nuclear Sciences, Lower Hutt.
Giggenbach, W. F., Y. Sano, and H. Wakita, 1993, Isotopic composition of He, and CO 2 and CH 4 contents in gases produced
along the New Zealand part of a convergent plate boundary:
Geochimica et Cosmochimica Acta, v. 57, p. 34273455.
Gould, R., 1980, The coal-forming flora of the Walloon coal measures: Coal Geology, v. 1, p. 83105.
Haskell, T., 1991, An analysis of Taranaki Basin hydrocarbon
migration paths; some questions answered: Petroleum Exploration New Zealand News, Jan., p. 1925.
Haskell, T., 1992, An analysis of Taranaki Basin hydrocarbon migration systems. Proceedings of 1991 New Zealand Oil Exploration
Conference, Ministry of Commerce, Wellington, p. 146.
Hedberg, H. D., 1968, Significance of high wax oils with respect to
genesis of petroleum: AAPG Bulletin, v. 52, p. 736750.
Hill, P. J., and J. D. Collen, 1978, The Kapuni sandstones from
Inglewood-1 well, Taranakipetrology and the effect of diagenesis on reservoir characteristics: New Zealand Journal of Geology and Geophysics, v. 21, p. 215228.
Hirner, A. V., and G. L. Lyon, 1989, Stable isotope geochemistry of
crude oils and of possible source rocks from New Zealand1:
carbon: Applied Geochemistry, v. 4, p. 109120.
Hirner, A. V., and B. W. Robinson, 1989, Stable isotope geochemistry of crude oils and of possible source rocks from New
Zealand2: sulphur: Applied Geochemistry, v. 4, p. 121130.
Holloway, P. J., 1982, Structure and histochemistry of plant cuticular membranes: an overview, in D. F. Cutler, K. L. Allen, and
C. E. Price, eds., The plant cuticle: Linnean Society, London,
Academic Press, p. 132.
Isaac, M. J., and R. H. Herzer, 1994, Beyond Taranakiis the next
Maui field west of Northland?: Proceedings of the New Zealand

Killops et al.

Petroleum Conference 1994, Ministry of Commerce, Wellington, p. 129.


Johnston, J. H., 1992, The characterisation of potential hydrocarbongenerative source rocks in the North Taranaki Basin by geochemical biomarkers: Proceedings of New Zealand Oil Exploration Conference 1991, Ministry of Commerce, Wellington, p. 357363.
Johnston, J. H., R. J. Collier, and J. D. Collen, 1990, What is the
source of Taranaki oils? Geochemical biomarkers suggest it is
the very deep coals and shales: Proceedings of New Zealand
Oil Exploration Conference 1989, Ministry of Commerce,
Wellington, p. 288295.
Johnston, J. H., R. J. Collier, and A. I. Maidment, 1991, Coal as
source rocks for hydrocarbon generation in the Taranaki Basin,
New Zealand: a geochemical biomarker study: Journal of
Southeast Asian Sciences, v. 5, p. 283289.
Kamp, P. J. J., and P. F. Green, 1990, Thermal and tectonic history
of selected Taranaki Basin (New Zealand) wells assessed by
apatite fission track history: AAPG Bulletin, v. 74, p. 14011419.
Khorasani, G. K., 1987, Oil-prone coals of the Walloon coal measures, Surat Basin, Australia, in A. C. Scott, ed., Coal and coalbearing strata: recent advances: Geological Society Special Publication, no. 32, Blackwell Scientific, Oxford, p. 303310.
Killops, S. D., and N. L. Frewin, in press, Triterpenoid diagenesis
and cuticular preservation: Organic Geochemistry.
Killops, S. D., and V. J. Howell, 1991, Complex series of pentacyclic triterpanes in a lacustrine sourced oil from Korea Bay
Basin: Chemical Geology, v. 91, p. 6579.
Killops, S. D., and V. J. Killops, 1993, An introduction to organic
geochemistry: Longman, Harlow.
King, P. R., 1990, Polyphase evolution of the Taranaki Basin, New
Zealand: changes in sedimentary and structural style: Proceedings of New Zealand Oil Exploration Conference 1989, Ministry of Commerce, Wellington, p. 134150.
King, P. R., and G. P. Thrasher, 1992, Post-Eocene development of
the Taranaki Basin, New Zealand: convergent overprint of a
passive margin, in J. S. Watkins, F. Zhiqiang and K. J. Miller,
eds., Geology and geophysics of continental margins: AAPG
Memoir 53, p. 93118.
King, P. R., T. R. Naish, and G. P. Thrasher, 1991, Structural cross
sections and selected palinspastic reconstructions of the
Taranaki Basin, New Zealand: New Zealand Geological Survey
Report G-150, Institute of Geological and Nuclear Sciences,
Lower Hutt.
Kirkland, D. W., T.-F. Tsui, and M. L. Stockton, 1987, Generation
of oil from coals and carbonaceous shales: AAPG Bulletin,
v. 71, p. 577.
Le Berre, F., S. Derenne, C. Largeau, J. Connan, and C. Berkaloff,
1991, Occurrence of non-hydrolysable, macromolecular, wall
constituents in bacteria, in D. A. C. Manning, ed., Organic geochemistry. Advances and applications in energy and the natural
environment: University Press, Manchester, p. 428431.
Lewan, M. D., 1984, Factors controlling the proportionality of
vanadium to nickel in crude oils: Geochimica et Cosmochimica
Acta, v. 48, p. 22312238.
Lowery, J. H., 1988, Catalogue of vitrinite reflectance measurements and coal analyses from oil prospecting wells in Taranaki
Basin, New Zealand: Report M168, Institute of Geological and
Nuclear Sciences, Lower Hutt.
Lu, S.-T., and I. R. Kaplan, 1992, Diterpanes, triterpanes, steranes
and aromatic hydrocarbons in natural bitumens and
pyrolysates from different humic coals: Geochimica et Cosmochimica Acta, v. 56, p. 27612788.
McBeath D. M., 1977, Gas-condensate fields of the Taranaki Basin,
New Zealand: New Zealand Journal of Geological Geophysics,
v. 20, p. 99127.
McKirdy D. M., and A. R. Chivas, 1992, Nonbiodegraded aromatic
condensate associated with volcanic supercritical carbon dioxide, Otway Basin: implications for primary migration from terrestrial organic matter: Organic Geochemistry, v. 18, p. 611627.
McKirdy, D. M., A. K. Aldridge, and P. J. M. Ypma, 1983, A geochemical comparison of some crude oils from pre-Ordovician

1583

carbonate rocks, in M. Bjory et al., eds., Advances in organic


geochemistry 1981: Wiley, Chichester, p. 99107.
McManamon, D., 1993, Kora exploration and drilling results:
Petroleum Exploration New Zealand News, July, p. 1827.
Mello, M. R., N. Telnaes, P. C. Gaglianone, M. I. Chicarelli, S. C.
Brassell, and J. R. Maxwell, 1988, Organic geochemical characterisation of depositional palaeoenvironments of source rocks
and oils in Brazilian marginal basins: Organic Geochemistry,
v. 13, p. 3145.
Mildenhall, D. C., 1980, New Zealand Late Cretaceous and Cenozoic plant biogeography: a contribution: Palaeogeography
Palaeoclimatology Palaeoecology, v. 31, p. 197233.
Mildenhall, D. C., 1988, Kaihu Group at Chases Gorge, in D. T.
Pocknall and R. Tremain, eds., New Zealand palynology and
palaeobotanya field guide to palynological and palaeobotanical localities: Institute of Geological and Nuclear Sciences,
Lower Hutt, p. 1215.
Moldowan, J. M., F. J. Fago, C. Y. Lee, S. R. Jacobson, D. S. Watt,
N.-E. Slougui, A. Jeganathan, and D. C. Young, 1990, Sedimentary 24-n-propylcholestanes, molecular fossils diagnostic of
marine algae: Science, v. 247, p. 309312.
Newman, J., and N. A. Newman, 1982, Reflectance anomalies in
Pike River coals: evidence of variability in vitrinite type, with
implications for maturation studies and Suggate rank: New
Zealand Journal of Geological Geophysics, v. 25, p. 233243.
Nip, M., E. W. Tegelaar, H. Brinkhuis, J. W. de Leeuw, P. A.
Schenck, and P. J. Holloway, 1986, Analysis of modern and fossil plant cuticles by Curie point Py-GC and Curie point Py-GCMS: Recognition of a new, highly aliphatic and resistant
biopolymer: Organic Geochemistry, v. 10, p. 769778.
Noble, R. A., R. Alexander, R. I. Kagi, and J. Knox, 1985, Tetracyclic diterpenoid hydrocarbons in some Australian coals, sediments and crude oils: Geochimica et Cosmochimica Acta,
v. 49, p. 21412147.
Noble, R. A., R. Alexander, R. I. Kagi, and J. Knox, 1986, Identification of some diterpenoid hydrocarbons in petroleum: Organic
Geochemistry, v. 49, p. 825829.
Noble, R. A., C. H. Wu, and C. D. Atkinson, 1991, Petroleum generation and migration from Talang Akar coals and shales offshore
N.W. Java, Indonesia: Organic Geochemistry, v. 17, p. 363374.
Palmer, J., and G. Bulte, 1991, Taranaki Basin, New Zealand, in K. T.
Biddle, ed., Active margin basins: AAPG Memoir 52, p. 261282.
Peakman, T. M., H. L. ten Haven, J. R. Rechka, J. W. de Leeuw, and
J. R. Maxwell, 1989, Occurrence of (20R)- and (20S)-8(14) and
14 5(H)-sterenes and the origin of 5(H),14(H),17(H)-steranes in an immature sediment: Geochimica et Cosmochimica
Acta, v. 53, p. 20012009.
Philp, R. P., and T. D. Gilbert, 1986, Biomarker distributions in
Australian oils predominantly derived from terrigenous source
material: Organic Geochemistry, v. 10, p. 7384.
Powell, T. G., 1988, Developments in concepts of hydrocarbon
generation from terrestrial organic matter, in H. C. Wagner,
L. C. Wagner, F. F. H. Wang, and F. L. Wong, eds., Petroleum
resources of China and related subjects: Earth Science Series,
v. 10, Circum-Pacific Council for Energy & Mineral Resources,
Houston, p. 807824.
Powell, T. G., C. J. Boreham, M. Smyth, N. Russell, and A. C. Cook,
1991, Petroleum source rock assessment in non-marine
sequences: pyrolysis and petrographic analysis of Australian
coals and carbonaceous shales: Organic Geochemistry, v. 17,
p. 375394.
Raine, J. I., C. P. Strong, and G. J. Wilson, 1993, Biostratigraphic
review of petroleum exploration wells, Great South Basin, New
Zealand: IGNS Sci. Rep. 93/32, Institute of Geological and
Nuclear Sciences, Lower Hutt.
Reed, J. D., 1992, Exploration geochemistry of the Taranaki Basin
with emphasis on Kora: Proceedings of the New Zealand
Exploration Conference 1991, Ministry of Commerce, Wellington, p. 364372.
Sandvik, E. I., W. A. Young, and D. J. Curry, 1992, Expulsion from
hydrocarbon sources: the role of organic absorption: Organic

1584

Oil Generation in the Taranaki Basin

Geochemistry, v. 19, p. 7787.


Sofer, Z, 1984, Stable carbon isotope compositions of crude oils:
application to source depositional environments and
petroleum alterations: AAPG Bulletin, v. 68, p. 3149.
Suggate, R. P., 1959, New Zealand coals: their geological setting
and its influence on their properties: Bulletin 134, Institute of
Geological and Nuclear Sciences, Lower Hutt.
Suggate, R. P., and J. P. Boudou, 1993, Coal rank and type variation
in Rock-Eval assessment of New Zealand coals: Journal of
Petroleum Geology, v. 16, p. 7388.
Summons, R. E., J. K. Volkman, and C. J. Boreham, 1987, Dinosterane and other steroidal hydrocarbons of dinoflagellate origin in
sediments and petroleum: Geochimica et Cosmochimica Acta,
v. 51, p. 30753082.
Sweeney, J. J., 1990, BASINMATFORTRAN program calculates oil and
gas generation using a distribution of discrete activation energies: Geobyte, v. 5, p. 3743.
Sweeney, J. J., and A. K. Burnham, 1990, Evaluation of a simple
model of vitrinite reflectance based on chemical kinetics:
AAPG Bulletin, v. 74, p. 15591570.
Sykes, R., R. P. Suggate, and P. R. King, 1992, Timing and depth of
maturation in Southern Taranaki Basin from reflectance and
rank(S): Proceedings of the New Zealand Oil Exploration Conference 1991, Ministry of Commerce, Wellington, p. 373389.
Tegelaar, E. W., J. W. de Leeuw, S. Derenne, and C. Largeau, 1989,
A reappraisal of kerogen formation: Geochimica et Cosmochimica Acta, v. 53, p. 31033106.
Thomas, B. M., 1982, Land-plant source rocks for oil and their significance in Australian basins: APEA Journal, v. 22, p. 164178.

Thrasher, G. P., 1990, The Maui field and the exploration potential
of Southern Taranaki; a few unanswered questions: Petroleum
Exploration New Zealand News, July, p. 2630.
Thrasher, G. P., 1992, Late Cretaceous source rocks of Taranaki
Basin: Proc. NZ Oil Explor. Conf. 1991, Ministry of Commerce,
Wellington, p. 147154.
van Aarssen, B. G. K., J. K. C. Hessels, O. A. Abbink, and J. W. de
Leeuw, 1992, The occurrence of polycyclic sesqui-, tri-, and
oligoterpenoids derived from a resinous polymeric cadinene in
crude oils from Southeast Asia: Geochimica et Cosmochimica
Acta, v. 56, p. 12311246.
Weston R. J., Z. Czochanska, C. M. Sheppard, A. D. Woolhouse,
and R. A. Cook, 1988a, Organic geochemistry of the sedimentary basins of New Zealand. Part III. A biomarker correlation of
three petroleum seeps and some rock bitumens from the East
Coast of the North Island: Journal of Royal Society of New
Zealand, v. 18, p. 225243.
Weston, R. J., M. H. Engel, R. P. Philp, and A. D. Woolhouse,
1988b, The stable carbon isotopic composition of oil from the
Taranaki Basin of New Zealand: Organic Geochemistry, v. 12,
p. 487493.
Weston, R. J., R. P. Philp, C. M. Sheppard, and A. D. Woolhouse,1989, Sesquiterpanes, diterpanes and other higher terpanes in oils from the Taranaki Basin of New Zealand: Organic
Geochemistry, v. 14, p. 405421.
Woolhouse, A.D., J.-N. Oung, R. P. Philp, and R. J. Weston, 1992,
Triterpanes and ring-A degraded triterpanes as biomarkers
characteristic of Tertiary oils derived from predominantly higher plant sources: Organic Geochemistry, v. 18, p. 2331.

ABOUT THE AUTHORS


Steve Killops
Steve Killops is an organic geochemist in the Hydrocarbon Resources and Basin Studies Group at
the Institute of Geological & Nuclear Sciences (IGNS). He has been
involved in petroleum geochemistry since obtaining a B.Sc. degree
and a Ph.D. at Bristol University.
Following a postdoctoral research
fellowship, Steve worked in the
commercial sector on biomarker
studies for the exploration industry and, prior to joining
IGNS at the end of 1992, he spent five years as lecturer
in organic geochemistry at Royal Holloway and Bedford
New College, University of London. He is the author of
An Introduction to Organic Geochemistry (Longman,
1993).

Tony Woolhouse
Tony Woolhouse is an organic
chemist with Industrial Research
Ltd. (IRL). He graduated in 1973
from the Victoria University of
Wellington (B.Sc. and Ph.D.) and,
following postdoctoral work
(19741975) at the University of
Liverpool, he joined the Organic
Chemistry Group of the Chemistry
Division of the then Department of
Scientific and Industrial Research
(now part of IRL). During 19841989, Tony set up a
research group to evaluate the genetic relationships of
oils and sediments in New Zealand from biomarker distributions, and also worked for a time with Paul Philp at
the University of Oklahoma.

Killops et al.

Rod Weston
Rod Weston is a graduate of the
Universities of Auckland (M.Sc.),
Oxford (D.Phil.), and London
(DIC). He joined the Chemistry
Division of the Department of Scientific and Industrial Research
(now Industrial Research Ltd.) in
1965 as an undergraduate and has
been with them ever since. He is
an organic chemist with interests
in the chemistry of natural products, especially essential oils, terpenoids, and steroids,
and has worked with Sir Derek W. R. Barton in London
and R. Paul Philp in Oklahoma.

1585

Richard Cook
Richard Cook received his Ph.D.
(Petroleum Geochemistry of
Taranaki Basin) from Victoria University of Wellington in 1987. He
has been part of the Basin Studies
Group of IGNS since joining what
was then the New Zealand Geological Survey (NZGS) in 1978. Richard
has published numerous papers on
basin studies, oil seeps, petroleum
geochemistry, and resource evaluation. He has also been business manager and group commercial coordinator in both DSIR and IGNS. Prior to
joining the NZGS, he worked for Texaco for four years
in exploration in the North Sea, Portugal, south Texas,
and international new ventures.

You might also like