Vous êtes sur la page 1sur 42

Diagenesis and reservoir quality of the Lower Cretaceous Quantou

Formation tight sandstones in the southern Songliao Basin, China

Kelai Xi, Yingchang Cao, Jens Jahren, Rukai Zhu, Knut Bjrlykke, Beyene
Girma Haile, Lijing Zheng, Helge Hellevang

PII: S0037-0738(15)00225-0
DOI: doi: 10.1016/j.sedgeo.2015.10.007
Reference: SEDGEO 4925

To appear in: Sedimentary Geology

Received date: 16 August 2015


Revised date: 20 October 2015
Accepted date: 23 October 2015

Please cite this article as: Xi, Kelai, Cao, Yingchang, Jahren, Jens, Zhu, Rukai, Bjrlykke,
Knut, Haile, Beyene Girma, Zheng, Lijing, Hellevang, Helge, Diagene-sis and reservoir quality
of the Lower Cretaceous Quantou Formation tight sand-stones in the southern Songliao Basin,
China, Sedimentary Geology (2015), doi: 10.1016/j.sedgeo.2015.10.007

This is a PDF file of an unedited manuscript that has been accepted for publication. As a
service to our customers we are providing this early version of the manuscript. The manuscript
will undergo copyediting, typesetting, and review of the resulting proof before it is published in
its final form. Please note that during the production process errors may be discovered which
could aect the content, and all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

Diagenesis and reservoir quality of the Lower Cretaceous

Quantou Formation tight sandstones in the southern Songliao

Basin, China

Kelai Xi a,b*, Yingchang Cao a*, Jens Jahren b, Rukai Zhu c, Knut Bjrlykke b, Beyene Girma Haile b, Lijing Zheng d, Helge

Hellevang b
a.School of Geosciences, China University of Petroleum, Qingdao, Shandong, 266580, China; b.Department of

Geosciences, University of Oslo, P.O. Box 1047, Blindern, 0316 Oslo, Norway; c.Research Institute of Petroleum

Exploration & Development, Beijing 100083, China; d. College of Energy, China University of Geosciences, Beijing,

100083, China

*Corresponding author:
E-mail: kelai06016202@163.com
xi.kelai@geo.uio.no

Abstract: The Lower Cretaceous Quantou Formation in the southern Songliao Basin
is the typical tight oil sandstone in China. For effective exploration, appraisal and
production from such a tight oil sandstone, the diagenesis and reservoir quality must
be thoroughly studied first. The tight oil sandstone has been examined by a variety of
methods, including cores and thin sections observation, XRD, SEM, CL, fluorescence,
electron probing analysis, fluid inclusion and isotope testing and quantitative
determination of reservoir properties. The sandstones are mostly lithic arkoses and
feldspathic litharenites with fine to medium grain size and moderate to good sorting.
The sandstones are dominated by feldspar, and quartz dominated and volcanic rock
fragments showing various stages of disintegration. The reservoir properties are quite
poor, with low porosity (average 8.54%) and permeability (average 0.493 mD), small
pore-throat radius (average 0.206 m) and high displacement pressure (mostly higher
than 1 MPa). The tight sandstone reservoirs have undergone significant diagenetic
alterations such as compaction, feldspar dissolution, quartz cementation, carbonate
cementation (mainly ferrocalcite and ankerite) and clay mineral alteration. As to the
onset time, the oil emplacement was prior to the carbonate cementation but posterior
to the quartz cementation and feldspar dissolution. The smectite to illite reaction and
pressure solution at stylolites provide a most important silica sources for quartz
cementation. Carbonate cements increase towards interbedded mudstones. Mechanical
compaction has played a more important role than cementation in destroying the
reservoir quality of the K1q4 sandstone reservoirs. Mixed-layer illite/smectite and illite
reduced the porosity and permeability significantly, while chlorite preserved the
porosity and permeability since it tends to be oil wet so that later carbonate
cementation can be inhibited to some extent. It is likely that the oil
1
ACCEPTED MANUSCRIPT

emplacement occurred later than the tight rock formation (with the porosity close to
10%). However, thicker sandstone bodies (more than 2 m) constitute potential
hydrocarbon reservoirs.
Key words: Tight sandstone diagenesis; reservoir quality; quartz cement; carbonate
cements; oil emplacement; Songliao Basin

1. Introduction
As one of the most important unconventional hydrocarbon resources, tight
sandstone oil is widely distributed in major petroliferous basins, potentially forming
large-scale petroleum reserves in China (Zou et al., 2013; Zou et al., 2014; Wang et al.,
2015). Tight sandstone is defined as reservoirs with a porosity less than 10%, in situ
formation permeability less than 0.1 mD or air permeability less than 1 mD in Chinese
basins (Zou et al., 2012). Reservoir quality is considered as the primary factor in tight
sandstone oil exploration (Zou et al., 2013; Fic and Pedersen, 2013; Stroker et al.,
2013). In general, tight sandstone reservoirs are deeply buried and have gone through
complicated diagenetic alterations, progressively changing the reservoir quality
(Vinchon et al., 1996; Karim, et al., 2010; Yang et al., 2012; Zhang, et al., 2015).
Therefore, it is important to have a detailed understanding of the diagenesis (Rahman
and McCann, 2012). However, there are still uncertainties related to quantification of
mineral dissolution and precipitation processes that have significant effects on
reservoir quality (Schmid et al., 2004; Gier et al., 2008; Taylor et al., 2010; Bjrlykke,
2014). Diagenetic evolution (time and temperature), cement sources and mass transfer
(open vs closed system) in sediments as well as their impacts on reservoir quality are
still debated (Schmid et al., 2004; Taylor et al., 2010; Bjrlykke and Jahren, 2012;
Yuan et al., 2015). In addition, the role of hydrocarbon emplacement on mineral
reactions is also still debated (Cao et al., 2012; Liu et al., 2014).
The Lower Cretaceous Quantou Formation tight sandstone is a prolific oil-
producing unit in the southern Songliao Basin (Li et al., 2013). Although there are
many publications dealing with stratigraphy, sedimentology and hydrocarbon
accumulation in the southern Songliao Basin, little attention has been paid to sandstone
diagenesis and reservoir quality evaluation (Zhang et al., 2007; Li et al., 2007; Hu et
al., 2008; Xiong et al., 2008; Feng et al., 2013; Dong et al., 2014). Authigenic quartz,
carbonate cements, dissolved feldspars and clay minerals are commonly observed in
the reservoirs determining the reservoir quality (Li et al., 2013; this study). Former
studies on diagenesis in this area only focused on quartz cement and its origin, but did
not investigate in detail any other diagenesis reactions linked to reservoir quality (Xi et
al., 2015). Furthermore, the origin of the carbonate cements and timing of diagenesis,
especially the diagenetic history and the processes
2
ACCEPTED MANUSCRIPT

controlling reservoir quality, have not been studied thoroughly. Understanding


quantitative diagenetic processes (including oil emplacement) and cement sources in
sandstones and their impact on reservoir quality are essential to further exploration,
appraisal and production of tight sandstone oil within this area.
The objectives of this paper were mainly focused on the different aspects of
diagenesis compared to the former studies focusing only on quartz cementation and
its origin: (1) perform a detailed diagenetic analysis and identify the sources of the
carbonate cements in these tight sandstones; (2) reconstruct the diagenetic history of
the tight sandstones and evaluate if oil emplacement affected the inorganic diagenesis;
(3) assess the effects of the different diagenetic processes in time and space on the
reservoir quality.

2. Geological background
The Songliao Basin is a Jurassic-Neogene lacustrine basin with a dual-structure
fault-depression in northeastern China (Fig. 1). The basin is located between 4225 to
4 2
4923 N and 11940 to 12824 E with an area about 2610 km (Zhang and Zhang,
2013). It can be further subdivided into seven first class tectonic zones (Zhou et al.,
2012), namely the Western Slope Zone, Northern Pitching Zone, Central Depression
Zone, Northeastern Uplift Zone, Southeastern Uplift Zone, Southwestern Uplift Zone
and Kailu Depression Zone (Xi et al., 2015) (Fig. 1). The study area, as one of the
most oil-rich areas, belongs to the Central Depression Zone and consists of three
secondary class tectonic units (Li et al., 2013), namely the Changling Sag, Huazijing
Terrace and Fuxin Uplift (Fig. 1).
Based on the filling sequence and structures, the basin evolution can be divided
into four phases: (1) a pre-rift phase, (2) a syn-rift phase, (3) a post-rift phase, and (4)
a compression phase (Zhang et al., 2009). Sediments filling the basin comprise the
Lower Cretaceous Huoshiling (K1h), Shahezi (K1sh), Yingcheng (K1yc), Doulouku
(K1d) and Quantou (K1q) Formations, the Upper Cretaceous Qingshankou (K 2qn),
Yaojia (K2y), Nenjiang (K2n), Sifangtai (K2s) and Mingshui (K2m) Formations, the
Palaeogene Yian (Ny) Formation, the Neogene Daan (Nd) and Kangtai (Nt)
Formations, and the Quaternary Pingyuan (Q) Formation (Xi et al., 2015) (Fig. 2).
Each formation can be further subdivided into different members (Fig. 2). The
Cretaceous stratigraphy contains many source rocks and reservoir rocks, which can
form different oil and gas accumulations vertically (Xi et al., 2015) (Fig. 2).
Previous studies indicated that the Songliao Basin has had a high geothermal
gradient throughout most of its history. From about 90.7 Ma to 65 Ma, the geothermal
o o
gradient has been estimated at 4.5-5.5 C/100 m, decreasing to about 4.0 C/100 m
after 65 Ma (Liu, 2004). During the whole burial process the Quantou formation has
been in a chemically closed system (Xi, et al., 2015). Presently, the sedimentary
3
ACCEPTED MANUSCRIPT

sequence is not at its maximum depth and temperature. The temperature of Quantou
o
Formation has never been more than about 130 C.
The studied section (Fig. 2), that is the fourth member of Quantou Formation
(K1q4), was deposited during the depression period of the tectonic evolution, and
mainly consists of delta sandstones and some interbedded mudstones (Xi et al., 2015).
According to oil analysis and source rock correlation, the first member of the
Qingshankou Formation (K2qn1), which just overlies K1q4, is the main source rock of
K1q4 reservoirs (Zou et al., 2005; Li et al., 2013). The oil generation (Ro = 0.5%) in
o
the K2qn1 began at a depth of about 1350-1450 m and at a temperature of 80-85 C
(Dong et al., 2014). The K1q4 sandstone reservoir is tight and strongly heterogeneous.
All wells contain oil, but with low production rates (Li et al., 2013). Reservoir quality
is considered as the most important control on hydrocarbon volumes in this area (Li et
al., 2013).

3. Databases and methods


The study was focused on the fourth member of the Quantou Formation in the
southern Songliao Basin where most of the producing oil fields are located.
Consequently, most of the drilling for oil has taken place in this area producing core
material, geophysical well logs and many other exploration and production well data
made available for this research. The data used in this paper was derived from more
than 40 wells (Fig. 1).
Rock composition data of 743 thin section samples, 8,529 reservoir porosity and
permeability data points, 622 grading analysis data, 275 mercury injection capillary
pressure testing data and well logging data for all the related wells, were obtained
from the Research Institute of Petroleum Exploration & Development of the Jilin
Oilfield Company, PetroChina.
According to the study objectives and constraints of the collected data, both
sandstone and interbedded mudstone samples were selected from the K 1q4 drill cores
of 26 wells. More than 300 polished thin sections and about 210 blue or red epoxy
resin-impregnated thin sections were prepared for the analysis of rock mineralogy,
diagenesis and visual pore characteristics. Thin sections were partly stained with
Alizarin Red S and K-ferricyanide for carbonate mineral identification. Point counting
was performed on 30 thin sections to check the correctness of the collected rock
composition data, which provided a standard deviation of 5% or less. For the content
of quartz cement, carbonate cements, primary pores and the feldspar dissolution pores,
20 micrographs each of 76 blue or red epoxy resin-impregnated thin sections were
taken using the Zeiss Axioscope A1 APOL digital transmission microscope. Then
cements and pores in each micrograph were identified under the microscope and
sketched on computer using CorelDRAW software, and the total area of cements and
4
ACCEPTED MANUSCRIPT

pores in every micrograph was obtained using Image-Pro Plus software. Finally, the
percentages of the cements and pores were calculated by taking the average of all
values in the 20 micrographs from each thin section. A total of 105 reservoir sandstone
samples and 40 interbedded mudstone samples were analyzed for whole-rock (bulk)
and clay fraction ( 2 m) mineralogy using XRD. Preparation, analysis and
interpretation procedures were modified from Moore and Reynolds (1997) and Hillier
(2003).
Thirty-one representative samples were viewed under a Quanta FEG 450 scanning
electron microscope (SEM) equipped with an energy dispersive X-ray spectrometer
(EDX). Cathode luminescence (CL) analyses of 16 typical samples were made using
an Olympus microscope equipped with a CL8200-MKS CL instrument. Thirty-seven
core samples from 12 wells were prepared as thick doubly polished thin sections for
fluorescent color observation, fluid inclusion petrographic analyses and
microthermometric measurements. The microthermometry of fluid inclusions was
studied using a petrographic microscope equipped with a Linkam THMSG 600
heating and cooling stage which enables temperatures of phase transitions in the range
of 180 to 500 C. Precision was 1 C for the homogenization temperature (T h) and
0.1 C for the final ice melting temperature, respectively.
Based on the petrological studies, 87 organic matter-free sandstone samples and 30
interbedded mudstone samples were chosen for carbon and oxygen stable isotope
analyses. These samples were analyzed using a Thermo-Finnigan MAT 253 isotope
ratio mass spectrometer. Precision was 0.08 for O and 0.06 for C. Carbon and
oxygen stable isotope data are presented in the notation relative to the Vienna
PeeDee Belemnite (V-PDB) standards.
Among all the analyses mentioned above, point counting, fluorescence observation
and fluid inclusion analysis were done in the Basin Analyses and Reservoir Geology
Key Laboratory of the China University of Petroleum. XRD, SEM, and CL were
performed in the Key Laboratory of Oil and Gas Reservoirs of PetroChina. Carbon
and oxygen stable isotope testing was done in the Analytical Laboratory of the CNNC
Beijing Research Institute of Uranium Geology.

4. Results
4.1. Reservoir lithologies
Petrographic investigation of the K1q4 tight sandstones shows that the detrital
components comprise 32.1-62.4% quartz (average 42.86%), 10.3-42.8% feldspars
(average 26%) and 12.4-47.6% rock fragments (average 31.14%), indicating that the
sandstones are mostly lithic arkoses and feldspathic litharenites (Fig. 3A). The
majority of the detrital quartz grains are monocrystalline. Detrital feldspars in these
sandstones are mostly plagioclase and altered K-feldspar. The rock fragments contain
5
ACCEPTED MANUSCRIPT

10.0-47.0% volcanic rocks with an average of 28.32%, 1.0-25.0% sedimentary rocks


with an average of 1.26%, and 1.0-17% metamorphic rocks with an average of 1.76%.
Most of the sandstones in main reservoir intervals generally do not contain much
detrital matrix.
According to grading analysis, the sandstones are fine to medium-grained (Fig. 3B).
Sorting ranges from moderately well to well sorted (Fig. 3C), and the roundness of the
detrital grains varies from sub-rounded to rounded. Grain contacts are dominated by
pointed to tangential contacts, as well as concavo-convex contacts. Microscopic
stylolites can be found also between detrital grains in the sandstones.
4.2. Reservoir properties
4.2.1 Porosity and permeability
As a whole, the reservoir properties in the K1q4 sandstones are quite poor (Fig. 4).
Helium porosity of core samples ranges from 1.7% to 20% (mainly 2.0% to 14.0%)
with an average of 8.54% (Fig. 4A). Horizontal permeability ranges from 0.01 mD to
44.5 mD (mainly less than 1 mD) with an average of 0.493 mD (Fig. 4B). Porosity
and permeability of the K1q4 tight sandstones decrease with an increase in burial
depth (Fig. 5). However, helium porosity and horizontal permeability show a poor
2
correlation relationship (R = 0.4789).
4.2.2 Pore types and pore-throat characteristics
Thin section and SEM observation revealed the presence of three types of pores in
the K1q4 tight sandstone reservoirs. Primary intergranular pores, triangular or
polygonal in shape, are affected by compaction generally (Fig. 5). Secondary pores
are mainly associated with feldspar dissolution, which can enlarge the intergranular
pores or form new intragranular pores (Fig. 5). The development of micro-pores is
chiefly associated with diagenetic clay minerals. Among them, primary intergranular
pores have better connectivity than secondary pores and micro-pores.
The ratio of the visual pores in thin sections varies as the burial depth increases. In
general, the relative content of the primary pore decreases from an average of about
80% to 50% with increasing burial depth (Fig. 5), while the relative content of
secondary pores increases from an average of about 20% to 50% (Fig. 5). This is
because compaction and cementation destroy primary intergranular porosity during
burial.
The pore-throat types for the tight sandstones mainly include contracted throats,
flake-like, curved, and thin pipe-like throats. Mercury injection capillary pressure
analyses show that the displacement pressure and the pressure corresponding to a 50%
Hg-saturation of the reservoirs have a positive correlation with permeability (Fig. 4C).
Tight sandstone reservoirs always have displacement pressures higher than 1 MPa and
the pressures correspond to a 50% Hg-saturation approximate 10 MPa, while the
conventional sandstone reservoirs are lower than 0.1 MPa and approximate 1 MPa,
6
ACCEPTED MANUSCRIPT

respectively (Fig. 4C). The average pore-throat radius ranges from 0.018 m to 1.776
m (mainly 0.1 m to 0.35 m) with an average of 0.206 m (Fig. 4D).
4.3. Diagenetic mineralogy
The K1q4 tight sandstones have undergone significant diagenetic modification,
including compaction, feldspar dissolution, quartz and carbonate cementation, clay
mineral alteration and some other minor cementation types.
4.3.1 Quartz cement
Authigenic quartz, occurring in two different types of morphologies, is generally
one of the most common cements in the studied reservoir rocks. The first type is
partial to complete syntaxial overgrowths, which are easy to discriminate from the
detrital grains due to the existence of some dust rims (Fig. 6A). The other quartz
cement type is pore-filling cement. This type is difficult to distinguish from detrital
quartz grains by the polarizing microscope (Fig. 6B). The CL images, however,
clearly display the boundaries between detrital quartz grains and authigenic quartz,
where the detrital grain is brightly luminescent and the authigenic quartz is dark non-
luminescent (Fig. 6C). The quartz cement in the tight sandstone reservoirs mainly
occurs as relatively large aggregates of microcrystalline or macrocrystalline, euhedral
authigenic quartz approximately 30-100 m in size (Fig. 6A-D). In addition,
authigenic quartz is always found together with clay minerals, such as kaolinite,
mixed-layer illite/smectite, illite and chlorite (Fig, 6D-F).
Quartz cement can be observed in most of the studied samples. The authigenic
quartz often occludes primary pores, significantly reducing the porosity and bridging
narrow pore-throats (Fig. 6A-D). Quantitative statistical data revealed that quartz
cement ranges from 1.55% to 8.38% with an average of 5.62% (Fig. 7). The quartz
cement shows a slightly increasing trend with increasing burial depth (Fig. 7).
4.3.2 Feldspar dissolution
Detrital K-feldspar grains have encountered partial to complete dissolution (Fig.
8A). The shallowest K1q4 sandstones (about 600-700 m) commonly contain secondary
pores, indicating that some feldspar dissolution may occur at a relatively early burial
stage. The content of feldspar dissolution in the studied reservoir ranges from 0.46%
to 4.72% with an average of 2.85% and shows no significant trend with increasing
burial depth below 1000 m (Fig. 7). Generally, feldspar dissolution is texturally
associated with kaolinite and what is now a small amount of microcrystalline
authigenic quartz at shallow burial depth (Fig. 8B), and filamentous and flaky illite at
deeper burial depth. In addition, some primary pores and secondary pores formed
from K-feldspars dissolution are partly to completely filled by authigenic albite (Fig.
8C), indicating albitization of K-feldspar.
4.3.3 Carbonate cements
As another volumetrically predominant cement type, carbonate cements mainly
7
ACCEPTED MANUSCRIPT

include ferrocalcite (Fig. 8D) and ankerite (Fig. 8E). They occur as scattered euhedral
rhombs and pore-filling blocky or mosaic aggregates. The CL micrographs of
carbonate cements show a homogeneous reddish orange luminescence color (Fig. 6F).
In thin sections, ferrocalcite and ankerite are mainly pore-filling cements around the
euhedral quartz overgrowth or partly replace the authigenic quartz (Fig. 8G-I).
Sometimes they precipitated in feldspar dissolution pores as well (Fig. 8I), suggesting
the conclusion that ferrocalcite and ankerite formed during late diagenesis.
Core analyses suggest that carbonate cements are not distributed uniformly in the
K1q4 tight sandstone reservoirs. The content of carbonate cement ranges from 1.97%
to 12.93% with an average of 5.72% and shows an insignificant trend with an increase
in burial depth (Fig. 7). Although carbonate cements can occur anywhere within the
reservoir intervals, they develop more intensely along the sandstone and mudstone
contact surface (Fig. 9). For example, the samples at depths of 2192.2 m, 2218 m in
Well Qian 223 and 1853.7 m in Well Rang 24, which are very close to the sandstone
and mudstone contact surface, have much more carbonate cements than the samples at
depths of 2212.35 m in Well Qian 223 and 1846.09 m and 1872.2 m in Well Rang 24,
which are in central part of the sandstone body (Fig. 9). That is a common
phenomenon in other wells in the study area as well. As a result, the reservoirs close
to the sandstone and mudstone contact surface always have a poorer quality with
lower porosity and permeability (Fig 9).
The carbon and oxygen isotope data may help to decipher the timing and sources of
materials for carbonate cementation (Irwin et al., 1977; Armitage et al., 2010). XRD
analysis data and rock thin sections indicate that the carbonate cements in the studied
samples are mainly pure ferrocalcite and ankerite, even though very few samples
13
contain mixed carbonate cements (Table. 1). The C value (V-PDB) ranges from
18
-12.2 to -0.9 with an average of -8.11, and the O value (V-PDB) ranges
from -22.9 to -18.2 with an average of -20.66 (Table. 1).
According to the thin section and fluorescence observation, aqueous inclusions
(two phases) are commonly present in authigenic quartz (Xi et al., 2015), but
extremely rare in the carbonate cements (only 5 inclusions found). Microthermometry
studies show that the homogenization temperature of the aqueous inclusions ranges
o o o
from 98 C to 108 C with an average of 101 C in carbonate cements (Table. 2).
4.3.4 Minor cements
Pyrite and feldspar overgrowths are two kinds of minor cements observed in K 1q4
tight sandstone reservoirs. Pyrite framboids fill intergranular pores (Fig. 10A) or
occurs over the carbonate cements, which probably formed during two stages, namely
early and late. Feldspar overgrowth occurs as rare, small discrete or coalesced
adularia-like crystals around detrital feldspar. The overgrowths are corroded by
ankerite and ferrocalcite (Fig. 10B), and thus may predate carbonate cementation.
8
ACCEPTED MANUSCRIPT

However, these two cements occur in less than 1% of the sandstones, and have
negligible effects on reservoir quality.
4.3.5 Clay minerals
Various types of clay minerals with different amounts and textural habits occur in
the K1q4 tight sandstone reservoirs, as revealed by XRD and SEM, including kaolinite,
mixed-layer illite/smectite, illite and chlorite. Minor amounts of kaolinte occur mainly
as booklets of vermicularly stacked pseudohexagonal crystals. Kaolinite is always
accompanied by minor amounts of microcrystalline authigenic quartz (Fig. 10C).
Mixed-layer illite/smectite occurs as honeycomb-textured, approximately 70%-95%
illite with an R=1 or R=3 Reichweite order (Fig. 10D). Fibrous and flaky illite is
mainly presents in primary pores and sometimes on grain surfaces and in feldspar
dissolution pores, locally bridging pore-throats (Fig. 10E). Needle and rosette shaped
chlorite found in primary pores (Fig. 10F) and on grain surfaces is also an important
clay mineral in the studied tight sandstone reservoirs.
Mixed-layer illite/smectite is the dominating clay mineral in K 1q4 tight sandstone
reservoirs, and all the clay minerals show trends with increasing burial depth (Fig. 11).
Small amounts of kaolinite mainly exist at depths shallower than about 1800 m, and
reduce rapidly below that depth, particularly, deeper than about 2000 m (Fig. 11),
o
where temperatures approach to 120 C. On the contrary, the percentage of illite in the
mixed-layer illite/smectite increases quickly deeper than about 1800 m and 2000 m
(Fig 11). The content of chlorite and mixed-layer illite/smectite shows an increasing
trend between 1800 m-2150 m and 2150 m-2450 m, respectively (Fig. 11). Mixed-
layer illite/smectite contains less than 85% illite with R=1 Reichweite order above
about 1800 m, and increases to more than 85% illite with R=3 Reichweite order below
1800 m (Fig. 11).
4.4 Oil presence characteristics
Based on the thin section evidence and fluorescence observations, oil is present in
most of the samples in the K1q4 tight sandstones (Fig. 12). The fluorescence colors of
oil and hydrocarbon inclusions are all primarily blue and white (Fig. 12), indicating a
relatively mature and light oil (Zhang, et al., 2009; Dong et al., 2014). The oil can be
observed in the studied reservoirs in the following locations: (1) in the primary
intergranular pores (Fig. 12A); (2) in the intergranular pores of the quartz cement (Fig.
12B); (3) in the narrow gaps left by carbonate cementation (Fig. 12C, D); (4) in the
cleavage of feldspar (Fig. 12E); (5) in the secondary pores of feldspar dissolution (Fig.
12F, G); (6) in the micropores of authigenic clay minerals (Fig. 12H); (7) in the
hydrocarbon inclusions along the fractures (Fig. 12I). However, there are few oil
traces that can be observed between quartz grains and overgrowths (Fig. 12B), and few
hydrocarbon inclusions are captured by quartz cements. These phenomena indicate
that oil emplacement mainly started to occur after the initial formation time
9
ACCEPTED MANUSCRIPT

of authigenic quartz, reducing possible inclusion formation sites and feldspar


dissolution.

5. Discussion
5.1. Source of quartz cement
The origin of quartz cement has been debated a long time with respect to external
versus internal silica sources within a given sandstones (Gier et al., 2008; Islam,
2009). Because the solubility of SiO2 (aq) is extremely low in relatively closed system
(Ronald and Edward, 1990; Bjrlykke and Jahren, 2012), together with constraints of
water volume and considerable heterogeneity in porosity and permeability, neither
advective flow, thermal convection or diffusion can explain long distance and massive
transfer of external SiO2 (aq) into sandstones (Bjrlykke et al., 1988; Bjrlykke, 2011;
Bjrlykke and Jahren, 2012). Thus, the internal sources, such as biogenic silica,
feldspar dissolution, unstable volcanic rock fragments, clay mineral diagenesis and
pressure solution of detrital grains, are possible silica sources for quartz cement
(Bjrlykke and Egeberg, 1993; Kim and Lee, 2004; Peltonen et al., 2009; Hyodo et al.,
2014). On the basis of these theories, the sources of quartz cement have already been
thoroughly studied in the K1q4 tight sandstone reservoirs (Xi et al., 2015). The results
o
showed that the quartz cement formed in a continuous process from about 60 C to
o
130 C (Fig. 13). Mass balance calculation indicated that the smectite- to -illite
reaction can provide 50-60% silica source for the quartz cementation in the K1q4 tight
sandstone reservoirs, which may explain the quartz cement formed from 60 to 100 C,
especially 70-90 C (Xi et al., 2015). Furthermore, the other part of silica source for
the quartz cementation was mainly provided by pressure solution (chemical
compaction) between detrital quartz, which can explain the quartz cement formed
above about 100 C (Xi et al., 2015). In addition, the K-feldspar and kaolinite to illite
reaction can provide only a small amount of silica source for the quartz cement formed
from about 120 C to 130 C (Xi et al., 2015). The internal supplied silica precipitate
within a closed system where the transport mechanism is diffusion (Xi et al., 2015).

5.2. Source of carbonate cements


Carbonate cements have many potential sources, including internal (e.g. locally
reprecipitated detrital carbonate grains or bioclasts.), external (from adjacent
mudstones or source rocks, etc), or mixed sources (Gier, et al., 2008; Dutton, 2010).
Sandstone petrology and provenance evidences show no detrital carbonate grains and
bioclasts existing in the reservoirs, suggesting that internal sources were probably
insignificant for the carbonate cement.
13
The relatively low negative C values suggest that decarboxylation of organic
matter in adjacent mudstones and source rocks must have been an important carbon
10
ACCEPTED MANUSCRIPT

source in the K1q4 tight sandstone reservoirs (Table. 1; Fig. 14A) (Irwin et al., 1977;
13
Curtis, 1987; Morad, 1998). The C values increase with the increasing distance of
samples to source rocks (Fig. 14B), also providing an evidence for carbonate cements
affected by organic matter decarboxylation from adjacent mudstones and source
rocks.
2+ 2+ 2+
In order to precipitate the ferrocalcite and ankerite, Ca , Mg and Fe ions must be
present in the formation water. As mentioned above, a large amount of unstable volcanic
rock fragments are present in the studied interval and smectite- to -illite reactions are very
common as well. Conversion of volcanic rock fragments and the smectite- to -illite
2+ 2+ 2+
reaction can provide extra Ca , Mg and Fe (Boles and Franks, 1979; Stroker et al.,
2013). As in the sandstones, the smectite- to -illite reaction can also occur in the
4+ 3+
interbedded mudstones and source rocks at the right temperature. Unlike Si and Al ,
2+ 2+ 2+
however, Ca , Mg and Fe can be expelled from mudstones to adjacent sandstones
(McHargue and Price, 1982; Chen et al., 2009; Yuan et al., 2015). Thus, their
concentrations along the sandstone and mudstone contact surface are higher than in the
central part of the sandstone body that has insignificant changes in lithological
composition. When these ions are released into the pore water of sandstones and mixed
2-
with the CO3 derived from organic matter decarboxylation, the initial physical and
chemical equilibrium is broken, allowing carbonate to precipitate (Milliken and Land,
1993; Chen et al., 2009; Dutton, 2010; Li et al., 2014). In this case, the carbonate cements
in sandstones may form along the sandstone and mudstone contact surfaces first, and then
gradually spread into the sandstone bodies by diffusion. The distribution patterns of
carbonate cements in sandstones, i.e. that extensive cementation occurs along the
sandstone and mudstone contact surfaces (Fig. 9), also supports external sources of
carbonate cements in the K1q4 tight sandstone reservoirs, which are related to interbedded
mudstones and source rocks.
Bulk rock XRD data show that interbedded mudstones contain an average of 2.82%
13
carbonate minerals. The C value (V-PDB) ranges from -13.3 to -0.9 with an
18
average of -6.17, and the O value (V-PDB) ranges from -21.7 to -12.1 with
an average of -16.46 (30 samples), which are similar to the carbonate cements in
sandstones (Table. 1). With relative high temperature, large amounts of organic acids
and CO2 derived from the thermal evolution of organic matter can dissolve some
carbonates in interbedded mudstones (Dutton, 2010), which may be another carbon
source of carbonate cements. In addition, plagioclase dissolution and associated
2+
illitization of kaolinite can also provide some Ca for carbonate cementation
(Macaulay et al., 1993; Milliken et al., 1994). However, these two sources are not
important since feldspar dissolution and carbonate minerals are very limited in the
studied reservoirs and interbedded mudstones, respectively.

11
ACCEPTED MANUSCRIPT

5.3. Paragenetic sequence of diagenesis


Below, petrographic observations will be integrated with fluid inclusion and isotope
analysis data to help discern the relative timing of each diagenesis and reconstruct the
diagenetic history and hydrocarbon emplacement process in the K 1q4 tight sandstone
reservoirs.
Petrographic evidence described in the preceding part of the text, including cross-
cutting relationships of cements, dissolution-filling relationships of diagenetic
minerals and fluorescent bitumen can be used to infer the relative sequence of each
diagenetic event. Moreover, the authigenic mineral forming temperatures that can be
measured from aqueous inclusions or calculated by oxygen isotope values are able to
provide a more accurate relative timing of the different diagenetic reactions. The
o
homogenization temperature of the aqueous inclusions ranges from 60 to 130 C (Fig.
o
13) with an average of 82.1 C in authigenic quartz (Xi et al., 2015).
In order to make up for the data deficiencies of the homogenization temperature in
18
carbonate cements, the O values are used to calculate the precipitation
13 18
temperatures of carbonate cements. The obtained C and O values cannot
18
represent the original pore water, so it is necessary to estimate the O values before
the calculation. For Jurassic and younger samples, the best discrimination between
marine and fresh-water carbonate is given by the equation of Keith and Weber (1964):
13 18
Z = a ( C+50) +b ( O+50), (3)
in which a and b are 2.048 and 0.498, respectively.
Marine carbonate is characterized by a Z-value above 120, fresh-water types with a
Z-value below 120, and indeterminate types with a Z-value near 120 (Keith and
Weber, 1964). The calculation results of the studied reservoirs show that the Z-values
range from 91.52 to 115.8 with an average of 100.75 (Table. 1), which indicate that the
origin of the pore water is mainly fresh-water with a slight mixed of salt-water. This
calculated result coincides with the final ice melting temperature and salinity data
18
measured from aqueous inclusions in carbonate cements (Table. 2). Using the O
o
values of these carbonate cements, the assumed precipitation temperature of 20 C and
the fractionation equation of Friedman and ONeil (1977), the oxygen isotope value of
the pore water from which carbonate cements have precipitated is approximately -7
18
relative to SMOW, which corresponds to the O value of pore water mixing of fresh
water and slightly salt water (Mansurbeg, et al., 2008). Finally, combined with the
18
O value of -7 (SMOW), the XRD analysis data and different oxygen isotope
fractionation factors for dolomite-water (Matthews and Katz, 1977) and calcite-water
(Friedman and O'Neil, 1977), the approximate carbonate cements precipitation
temperatures are calculated for all the studied samples. The calculated temperature
o o o
range from 83.78 C to 130.96 C with an average of 108.59 C (Table. 1), which is
similar to the homogenization temperature measured from carbonate cements
12
ACCEPTED MANUSCRIPT

(Table. 2).
There are many hydrocarbon inclusions developed in quartz micro-fractures (Fig.
12I). The coeval aqueous inclusion homogenization temperature is a good
approximation of the trapping temperature of the coexisting hydrocarbon inclusions
(Dong, et al., 2014). Microthermometry studies show that the homogenization
o o
temperature of the coeval aqueous inclusions ranges from 73 C to 119 C, with an
o
average of 93.08 C (Fig. 15). By comparison based on this result, the start of oil
emplacement was prior to the onset of carbonate cementation but posterior to the
onset of quartz cementation (Fig. 15).
In summary, synthesizing the petrographic evidences, the mineral reactions
associated with cement source analysis, and the calculated temperature of major
cements and oil emplacement, the paragenetic sequence of diagenesis and the
hydrocarbon emplacement are reconstructed and illustrated in Figure 16.
5.4. Diagenetic controls on reservoir quality
Reservoir quality, defined here as porosity and permeability, is a function of
provenance and depositional controls on grain size and sorting as well as diagenetic
controls of compaction, cementation, dissolution and development of authigenic clay
minerals (Schmid et al., 2004; Islam, 2009; Rahman and McCann, 2102). In the
studied reservoirs, the grain size and sorting are slightly different in each sample,
probably with insignificant effects on reservoir quality. Thus, the reservoir quality at
this stage is controlled mainly by diagenesis.
Mechanical compaction, starting immediately after deposition, can be evaluated by
the intergranular volume (IGV). The intergranular volume (IGV) found in the studied
samples from the petrographic analysis shows very little depth dependence, with
values between about 11.09% and 25.94%, with an average of 15.46% (Fig. 7). These
large variations in intergranular volume (up to 10%-15%) are probably related to both
the original volcanic content and the content of ductile volcanic grains. Sandstones
with abundant, mechanically unstable minerals commonly suffer more rapid reduction
in porosity and permeability during burial because of mechanical compaction. The
sandstones with more quartz grains will have a larger intergranular volume (Fig. 17);
however, the more feldspar grains and ductile rock fragments, the smaller the
intergranular volume will be (Fig. 17). This is because the quartz grains are hard and
more compression resistant. The rock fragments deform easily due to breakage along
grain boundaries resulting in ductile deformation lowering the IGV in-between the
fragmented lithoclasts and also between the disintegrated rock fragments and the
larger feldspar and quartz grains. Feldspar grains may also deform preferentially
during early burial due to weakening by chemical dissolution processes. Thus,
reservoirs with originally more quartz grains and less ductile rock fragment and
feldspar grains will be more mechanically more stable and preserve a higher IGV.
13
ACCEPTED MANUSCRIPT

A plot of intergranular volume (IGV) versus cement volume indicates that


mechanical compaction has played a more important role than cementation in
destroying the primary porosity of the K1q4 sandstone reservoirs (Fig. 17).
Mechanical compaction has accounted for 45.89%-66.51% (with an average of
58.08%) of the total intergranular pore volume loss (PVL) during the burial process
(assuming an initial porosity of 40%).
Cementation also played a major role in reducing porosity and permeability in the
studied sandstone reservoirs, even though mechanical compaction is volumetrically
the most important. The most abundant pore-occluding cement types recognized in
this study are quartz, carbonate and authigenic clay minerals. The different types of
cement occlude both pores and pore-throats, making the reservoir quality extremely
poor. Quartz cements were found in all samples. The amounts of quartz cement show
no significant trend with the distance to the sandstone and mudstone contact surface
(Fig. 19), indicating that quartz cement precipitation was not related to interbedded
mudstones. This could be regarded as evidence showing that the sources of quartz
cement were internal. Porosity destroyed by quartz cement ranges from 1.55% to
8.38% with an average of 5.62%.
Carbonate cements, formed relatively late during diagenesis based on isotopes, were
mainly related to adjacent mudstones or source rocks thicker than 2 m.
Quantitative analysis indicates that the carbonate cement content, porosity and
permeability were significantly influenced by proximity to the sandstone - mudstone
interface (Fig. 19). Generally, sandstones less than 1.0 m from the sandstone -
mudstone contact surface have a carbonate cement content above 8%, with porosity
and permeability lower than 6.5% and 0.05 mD respectively (Fig 19). On the other
hand, samples taken more than 7 m away from sandstone - mudstone contact surfaces
have carbonate cement contents less than 3%, and porosity and permeability higher
than 10% and 0.3 mD respectively (Fig. 19). The carbonate cement content ranges
from about 3% to 8%, and the porosity and permeability range from 6.5% to 10% and
0.05 mD to 0.3 mD, respectively (Fig. 19). Thicker sandstone bodies always show
enhanced cementation along the contact surfaces between sandstone and mudstone
intervals while retaining better reservoir properties relatively farther away from the
contact surface. However, thin sandstone layers (mainly less than 2 m) are evenly
carbonate cemented, resulting in poor reservoir properties in such thin sandstone
sections. The uneven distribution of carbonate cement found within the sandstones
indicates the intercalated mudstones as an external source of the carbonate cements.
Since the cement distribution and reservoir properties are a function of distance to the
interbedded mudstones (above or below) on a local scale, the system seems to have
been relatively closed (diffusion only) also with respect to late carbonate cementation.
According to the XRD analysis data, the total content of clay minerals ranges from
14
ACCEPTED MANUSCRIPT

4.1% to 26.5% with an average of 14.3%. The clay minerals present in the studied
sandstone reservoirs are mainly authigenic clays, consisting mostly of mixed-layer
illite/smectite with R=3 Reichweite order, illite, chlorite and a small amount of
kaolinite. The clay minerals reduced the porosity and permeability, occurring as pore-
lining and pore-filling clays. The relative abundance of the different clay minerals has
various effects on the reservoir qualities. Honeycomb mixed-layer illite/smectite
mainly occluded the primary pores and thus destroyed the reservoir quality, especially
the porosity (Fig. 20). Illite tended to plug the pores and pore-throats with fibrous and
flaky crystallite types, which can divide macro-pores into micro-pores and lower the
reservoir quality, particularly the permeability (Fig. 20). There is no evidence to
indicate that chlorite inhibited quartz cementation effectively, indicating that the
chlorite precursor was not attached to detrital mineral surfaces (quartz and feldspar).
However, there is a positive correlation between the content of chlorite and reservoir
properties (Fig. 20). Although chlorite can fill pores and reduce the diameter of pore-
throats, the crystals of chlorite are commonly small and do not cause severe occlusion
of pores (Morad et al., 2010). In addition, chlorite tends to be oil wet while illite is
commonly water wet (Barclay and Worden, 2000; Morad et al., 2010). Thus oil is
easier to accumulate in reservoirs with more chlorite. As a result, late carbonate
cementation may be inhibited to some extent due to mixed wettability (Fig. 12C, D).
The data from Well Rang 59 show that oil saturation increases with an increase in
chlorite, while carbonate has a decreasing trend (Fig. 21). Authigenic clay minerals
always contain a large amount of micropores that have negligible contributions for
permeability, which can be regarded as the most probable explanation for the
extremely low permeability and the poor relationship between porosity and
permeability in K1q4 sandstone reservoirs (Fig. 4A, B).
In addition, early feldspar dissolution resulted in the formation of some secondary
pores in the K1q4 sandstone reservoirs (Fig. 8A-C). Due to the formation of kaolinite
and weakening of compression resistance, however, feldspar dissolution does not
improve the reservoir quality effectively. Instead, the connectivity of secondary pores
is poor and thus makes the pore-throat texture of tight sandstones more intricate,
which is not an advantage with respect to oil accumulation and development.
The porosity and permeability will change due to diagenetic modifications. If one
assumes that the initial porosity of the studied reservoir was 40% (Houseknecht, 1987;
Lundegard, 1992), combing the diagenetic history and quantitative data of each
diagenetic modification, the approximate trend of the porosity (visual porosity from
the average of each thin section) for the K1q4 tight sandstone reservoirs can be seen in
Fig. 16. From this trend, compaction and quartz cementation are the major factors that
decreased the sandstone porosity. Carbonate cementation also increased the degree of
densification during later diagenesis. Moreover, pore-filling and pore-lining
15
ACCEPTED MANUSCRIPT

authigenic clay minerals, mainly formed at earlier stages during diagensis, also
damaged the reservoir quality significantly. As mentioned above, the oil emplacement
occurred mainly before carbonate cementation and after the onset of quartz
cementation. Most probably the oil emplacement occurred after most of the
sandstones had become tight reservoirs (with porosity close to 10%). This may help to
explain why most of the tight reservoirs have low oil saturation (mostly between
30%-40%) and that the oiliness is not dependent of porosity and permeability in the
K1q4 tight sandstone reservoirs.

6 Conclusions
1. The sandstones of K1q4 are mostly lithic arkoses and feldspathic litharenites with
a fine to medium grain size and moderate to good sorting, which are characterized by
abundant volcanic rock fragments.
2. The reservoir properties in the K1q4 sandstones are quite poor with low porosity
and permeability, small pore-throat radii and high displacement pressures. The relative
contents of the primary pores decreased with increasing burial depth, while the
amount of secondary pores shows an increasing trend.
3. The tight sandstone reservoirs have undergone significant diagenetic alterations
such as mechanical compaction, feldspar dissolution, quartz cementation, carbonate
cementation (mainly ferrocalcite and ankerite) and clay mineral alteration. Oil
emplacement was prior to the carbonate cementation but posterior to feldspar
dissolution and the onset of quartz cementation.
4. Smectite- to -illite reaction and pressure solution at stylolites provided the most
important internal silica sources for the quartz cementation; however, carbonate
cements mainly precipitated from external sources related to interbedded mudstones
and source rocks.
5. Mechanical compaction has played a more important role than cementation in
destroying the reservoir quality of the K1q4 sandstone reservoirs. The large variations
in intergranular volume, reflecting mechanical compaction, are mainly related to the
original volcanic content and the content of ductile volcanic grains.
6. Different chemical diagenetic processes have different impacts on reservoir
quality. Quartz cement is evenly distributed within the sandstone bodies, while
carbonate cement amounts are always higher along sandstone - mudstone contact
surfaces compared to the center of individual sandstone bodies. Due to the formation
of kaolinite and weakening of compression resistance, however, feldspar dissolution
has not improved the reservoir quality effectively.
7. The relative abundance of different clay minerals has various effects on the
reservoir qualities. Mixed-layer illite/smectite and illite reduced the porosity and
permeability significantly, while chlorite may have preserved the porosity and
16
ACCEPTED MANUSCRIPT

permeability somewhat since it tends to be oil wet and the mixed wettability may
have inhibited carbonate cementation to some extent.
8. Based on the evidence presented in this study, oil emplacement occurred first
when most of the sandstones had already reached tight reservoir conditions (porosity
close to 10%). However it is likely that thicker sandstone bodies within the studied
section (at least thicker than 2 m) could be potential hydrocarbon reservoirs.

7 Acknowledgments
The research was co-funded by the National Natural Science Foundation of China
(Grant No.U1262203), the Fundamental Research Funds for the Central Universities
(Grant No.14CX06013A, ), and the Chinese Scholarship Council (No.201406450019).
Engineers Desheng Zhu, Jianchang You and Jianming Li at the Research Institute of
Petroleum Exploration & Development are thanked for SEM, XRD and QEMSEM
analysis. We also thank the Jilin Oil field, PetroChina, for providing data.

References
Armitage, J.P., Worden, H.R., Faulkner, R.D., Aplin, C.A., Butcher, R.A., IliffeJ.,
2010. Diagenetic and sedimentary controls on porosity in Lower Carboniferous
fine-grained lithologies, Krechba field, Algeria: A petrological study of a
caprock to a carbon capture site. Marine and Petroleum Geology 27, 1395-1410.
Barclay, S.A., Worden, R.H., 2000. Effects of reservoir wettability on quartz
cementation in oil field. In: Worden, R.H., Morad, S. (Eds.), Clay Cements in
Sandstones, IAS Special Publication 29, 103-117.
Bjrlykke, K., 2011. Open-system chemical behaviour of Wilcox Group mudstones.
How is large scale mass transfer at great burial depth in sedimentary basins
possible? A discussion. Marine and Petroleum Geology 28, 1381-1382.
Bjrlykke, K., 2014. Relationships between depositional environments, burial history
and rock properties. Some principal aspects of diagenetic process in sedimentary
basins. Sedimentary Geology 301, 1-14.
Bjrlykke, K., Egeberg, P.K., 1993. Quartz cementation in sedimentary basins. AAPG
Bulletin 77, 1538-1548.
Bjrlykke, K., Jahren, J., 2012. Open or closed geochemical systems during
diagenesis in sedimentary basins: Constraints on mass transfer during diagenesis
and the prediction of porosity in sandstone and carbonate reservoirs. AAPG
Bulletin 96, 2193-2214.
Bjrlykke, K., Mo, A., & Palm, E., 1988. Modelling of thermal convection in
sedimentary basins and its relevance to diagenetic reactions. Marine and
Petroleum Geology 5, 338-351.
Bodnar, R.J., 1993. Reviced equation and table for determining the freezing point
17
ACCEPTED MANUSCRIPT

depression of H2O-NaCl solutions. Geochimica et Cosmochimica Acta 57, 683-


684.
Boles, J.R., Franks, S.G., 1979. Clay diagenesis in Wilcox sandstones of southwest
Texas: implications of smectite diagenesis on sandstone cementation. Journal of
Sedimentary Petrology 49, 55-70.
Cao, Y., Yuan, G., Wang, Y., Xi, K., Kuang, L., Wang, X., 2012. Genetic mechanisms
of low permeability of reservoirs of Qingshuihe Formation in Baisantai area,
Junggar Basin. Acta Petrolei Sinica 33, 758-771.
Chen, D., Pang, X., Jiang, Z., Zeng, J., Qiu, N., Li, M., 2009. Reservoir characteristics
and their effects on hydrocarbon accumulation in lacustrine turbidites in the
Jiyang super-depression, Bohai Bay Basin, China. Marine and Petroleum
Geology 26, 149-162.
Curtis, C.D., 1978. Possible links between sandstone diagenesis and depth-related
geochemical reactions occurring in enclosing mudstones, Journal of the
Geological Society 135, 107-117.
Dong, T., He, S., Wang, D., Hou, Y., 2014. Hydrocarbon migration and accumulation
in the Upper Cretaceous Qingshankou Formation, Changling Sag, southern
Songliao Basin: Insights from integrated analyses of fluid inclusion, oil source
correlation and basin modelling. Journal of Asian Earth Sciences 90, 77-87.
Dutton, P.S., Loucks, G.R., 2010. Diagenetic controls on evolution of porosity and
permeability in lower Tertiary Wilcox sandstones from shallow to ultradeep
(200-6700m) burial, Gulf of Mexico Basin, U.S.A. Marine and Petroleum
Geology 27, 69-81.
Feng, C., Shan, Q., Shi, W., Zhu, S., 2013. Reservoirs heterogeneity and its control on
remaining oil distribution of K1q4, Fuyu Oilfield. Joural of China University of
Petroleum 37, 1-7.
Fic, J., Pedersen, K.P., 2013. Reservoir characterization of a tight oil reservoir,
the middle Jurassic Upper Shaunavon Member in the Whitemud and Eastbrook
pools, SW Saskatchewan. Marine and Petroleum Geology 44, 41-59.
Friedman, N., ONeil, J.R., 1977. Compilation of stable isotope fractionation factor.
US Geological Survey Professional Paper 440-KK.
Gier, S., Worden, H.R., Johns, D.W., Kurzweil, H. 2008. Diagenesis and reservoir
quality of Miocene sandstones in the Vianna Austria. Marine and Petroleum
Geology 25, 681-695.
Guo, H., Wang, D., 1999. Stable isotopic composition and origin analysis of the
carbonate cements within sandstone reservoirs of T arim oil-gas bearing area.
Petroleum exploration and development 26, 31-32.
Hillier, S., 2003. Quantitative analysis of clay and other minerals in sandstones by X-
ray powder diffraction (XRPD). In: Worden, R., Morad, S. (Eds.), Clay
18
ACCEPTED MANUSCRIPT

Mineral Cements in Sandstones. nternational Association of Sedimentologist,


Special Publication. , Oxford, International 213-251.
Houseknecht, W.D., 1987. Assessing the Relative Importance of Compaction
Processes and Cementation to Reduction of Porosity in Sandstones. AAPG
Bulletin 71, 633-642.
Hu, X., Bao, Z., Na, W., Zhou, X., 2008. Sedimentary facies of the fourth member of
Quantou Formation in Fuyu oilfield , the South Songliao Basin . Oil and gas
geology 29, 334-341.
Hyodo, A., Kozdon, K., Pollington, D.A., Valley, W.J., 2014. Evolution of quartz
cementation and burial history of the Eau Claire Formation based on in situ
oxygen isotope analysis of quartz overgrowths. Chemical Geology 384, 168-180.
Irwin, H., Curtis, C., Coleman, M., 1977. Isotopic evidence for source of diagenetic
carbonates formed during bruial of organic-rich sediments . Nature 269, 209-
213.
Islam, A.M., 2009. Diagenesis and reservoir quality of Bhuban sandstones (Neogene),
Titas Gas Field, Bengal Basin, Bangladesh. Journal of Asian Earth Sciences 35,
89-100.
Karim, A., Piper, P.G., Piper, J.M., D., 2010. Controls on diagenesis of Lower
Cretaceous reservoir sandstones in the western Sable Subbasin, offshore Nova
Scotia. Sedimentary Geology 224, 65-83.
Keith, L.M., Weber, N.J., 1964. Carbon and oxygen isotopic composition of selected
limestones and fossils. Geochimica et. Cosmochimica Acta 28, 1787-1816.
Kim, Y., Lee, Y., 2004. Origin of quartz cement in the Lower Ordovician Dongjeom
formation, Korea. Journal of Asian Earth Sciences 24, 327-335.
Li, D., Dong, C., Lin, C., Ren, L., Jiang, T., Tang, Z., 2013. Control factors on tight
sandstone reservoirs below source rocks in the Rangzijing slope zone of southern
Songliao Basin, East China . Petroleum exploration and development 40, 692-
700.
Li, Q., Jiang, Z., Liu, K., Zhang, C., You, X., 2014. Factors controlling reservoir
properties and hydrocarbon accumulation of lacustrine deep-water turbidites in
the Huimin Depression, Bohai Bay Basin, East China. Marine and Petroleum
Geology 57, 327-344.
Liu, J., 2004. The simulation of thermal history in Changling Sag. Natural Gas
Industry 24, 17-20.
Liu, M., Liu, Z., Liu, J., Zhu, W., Huang, Y., Yao, X., 2014. Coupling relationship
between sandstone reservoir densification and hydrocarbon accumulation: A case
from the Yanchang Formation of the Xifeng and Ansai areas, Ordos Basin.
Petroleum Exploration and Development 41, 185-192.
Li, Y., Yu, K., Jiang, Y., Chen, S., Zhang, E., Song, Y., 2007. A New Explanation for
19
ACCEPTED MANUSCRIPT

the Stratigraphical Sequence of Quan 4th Member of the Fuyu Oil Layer in the
Songliao Basin. Periodical of Ocean University of China 37, 977-982.
Lundegard, P., 1992. Sandstone porosity loss - a big picture view of the
importance of compaction. Jounral of Sedimentary Petrology 62, 250-260.
Macaulay, C.I., Haszeldine, R.S., Fallick, A.E., 1993. Distribution, chemistry, isotopic
composition and origin of diagenetic carbonates:Magnus Sandstone, North Sea.
Journal of Sedimentary Research 63, 33-43.
Mansurbeg, H., Morad, S., Salem, A., Marfil, R., El-ghali, A.K.M., Nystuen, P.J.,
2008. Diagenesis and reservoir quality evolution of palaeocene deep-water,
marine sandstones, the Shetland-Faroes Basin, British continental shelf. Marine
and Petroleum Geology 25, 514-543.
Matthews, A., Katz, A., 1977. Oxygen isotope fractionation during the dolomitization.
Geochimica et. Cosmochimica Acta 41, 1431-1438.
McHargue, T.R., Price, R.C., 1982. Dolomite from clay in argillaceous limestones or
shale associated marine carbonates. Journal of Sedimentary Petrology 52, 873-
886.
Milliken, K.L., Land, L.S., 1993. The origin and fate of silt sized carbonate in
subsurface MioceneeOligocene mudstones, south Texas Gulf Coast.
Sedimentology 40, 107-124.
Milliken, K.L., Mack, L.E., Land, L.S., 1994. Elemental mobility in sandstones
Duping burial: whole-rock chemical and isotopic data, Frio Formation, South
Texas. Journal of Sedimentary Research 64, 788-796.
Moore, D.M., Reynolds, R.C., 1997. X-ray Diffraction and the Identification and
Analysis of Clay Minerals. Oxford: Oxford University Press.
Morad, S., 1998. Carbonate cementation in sandstones: distribution patterns and
geochemical evolution. International Association of Sedimentologists Special
Publication 26, 1-26.
Morad, S., Al-Ramadan, K., Ketzer, J.M., De Ro, L.F., 2010. The impact of diagenesis
on the heterogeneity of sandstone reservoirs: a review of the role of depositional
facies and depositional facies and sequence stratigraphy. AAPG Bulletin 94,
1267-1309.
Peltonen, C., Marcussen, ., Bjrlykke, K., Jahren, J., 2009. Clay mineral diagenesis
and quartz cementation in mudstones: The effects of smectite to illite reaction on
rock properties. Marine and Petroleum Geology 26, 887-898.
Rahman, J.J., M., McCann, T., 2012. Diagenetic history of the Surma Group
sandstones (Miocene) in the Surma Basin, Bangladesh. Journal of Asian Earth
Sciences 45, 65-78.
Ronald, K.S., Edward, D.P., 1990. Secondary porosity revisited: The chemistry of
feldspar dissolution by carboxylic: Acids and Anions. AAPG Bulletin 74,
20
ACCEPTED MANUSCRIPT

1795-1808.
Schmid, S., Worden, H.R., Fisher, J.Q., 2004. Diagenesis and reservoir quality of the
Sherwood Sandstone (Triassic), Corrib Field, Slyne Basin, west of Ireland.
Marine and Petroleum Geology 21, 299-315.
Storker, M.T., Harris, B.N., Elliott, C.W., Wampler, M.J., 2013. Diagenesis of a tight
gas sand reservoir: Upper Cretaceous Mesaverde Group, Piceance Basin,
Colorado. Marine and Petroleum Geology. 40, 48-68.
Taylor, T.R., Giles, M.R., Hathon, L.A., Diggs, T.N., Braunsdorf, N.R., Birbiglia, G.V.,
2010. Sandstone diagenesis and reservoir quality prediction: models, myths, and
reality. AAPG Bulletin 94, 1093-1132.
Vinchon, C., Giot, D., Sperber, O.F., Arbey, F., Thibieroz, J., Cros, P., 1996. Changes
in reservoir quality determined from the diagenetic evolution of Triassic and
Lower Lias sedimentary successions (Balazuc borehole, Ard che, France).
Marine and Petroleum Geology 13, 685-694.
Wang, J., Feng, L., Steve, M., Tang, X., Gail, E.T., Mikael, H., 2015. China's
unconventional oil: A review of its resources and outlook for long-term
production. Energy 82, 31-42.
Xi, K., Cao, Y., Jaren, J., Zhu, R., Bjrlykke, K., Zhang, X., Cai, L., Hellevang, H.,
2015. Quartz cement and its origin in tight sandstone reservoirs of the
Cretaceous Quantou formation in the southern Songliao basin, China. Marine
and Petroleum Geology 66, 748-763.
Xiong, F., Shi, G., Sun, Y., Sun, Y., 2008. Research on Sedimentary Microfacies of
Fourth Section of Quantou Formation in Liangjing Gudian Area of Songliao
Basin. Acta Scientiarum Naturalium Universitatis Pekinensis 44, 185-192.
Yang, R., Fan, A., Han, Z., Wang, X., 2012. Diagenesis and porosity evolution of
sandstone reservoirs in the East II part of Sulige gas field, Ordos Basin.
International Journal of Mining Science and Technology 22, 311-316.
Yuan, G., Gluyas, J., Cao, Y., Oxtoby, H.N., Jia, Z., Wang, Y., 2015. Diagenesis and
reservoir quality evolution of the Eocene sandstones in the northern Dongying
Sag, Bohai Bay Basin, East China. Marine and Petroleum Geology 22, 77-89.
Zhang, G., Bao, Z., Na, M., Zhou, X., Sun, J., 2007. Sedimentary facies of the
Member 4 of Quantou Formation of Lower Cretaceous in Southern Central
Depression of Songliao Basin . Journal of Palaeogeography 9, 267-276.
Zhang, L., Bai, G., Luo, X., Ma, X., Chen, M., Wu, M., 2009. Diagenetic history of
tight sandstones and gas entrapment in the Yulin Gas Field in the central area of
the Ordos Basin, China. Marine and Petroleum Geology 26, 974-989.
Zhang, W., Li, Y., Xu, T., Cheng, H., Zheng, Y., Xiong, P., 2009. Long-term variations of
CO2 trapped in different mechanismsin deep saline formations: A case study of the
Songliao Basin, China. International Journal of Greenhouse Gas Control 3,
21
ACCEPTED MANUSCRIPT

161-180.
Zhang, X., Zhang, L., 2013. Quantitive research on effective source rocks of the
Denglouku Formation in northern Songliao Basin. Chinese Journal of Geology
48, 879-890.
Zhang, Y., Piper, P.G., Piper, J.W., D., 2015. How sandstone porosity and permeability
vary with diagenetic minerals in the Scotian Basin, offshore eastern Canada:
Implications for reservoir quality. Marine and Petroleum Geology 63, 28-45.
Zhou, Z., Feng, D., Wang, H., Zhou, H., Wang, Q. 2012. Sequence stratigraphy
framework of Qing 1 and Qing 2 members in the western slope of southern
Songliao Basin and its control effect on reservoir sand bodies. Lithologic
reservoirs 24, 39-50.
Zou, C., Jia, C., Zhao, W., Tao, S., Gu, Z., Hou, J., 2005. Accumulation dynamics and
distribution of litho stratigraphic reservoirs in South Songliao Basin. Petroleum
exploration and development 32, 125-130.
Zou, C., Yang, Z., Zhang, G., Hou, L., Zhu, R., Tao, S., 2014. Conventional and
unconventional petroleum orderly accumulation: Concept and practical
significance. Petroleum Exploration and Develpoment 41, 14-30.
Zou,C., Zhang, G., Yang, Z., Tao, S., Hou, L., Zhu, R., 2013. Concepts,
characteristics, potential and technology of unconventional hydrocarbons: On
unconventional petroleum geology. Petroleum Exploration and Develpoment 40,
413-428.
Zou, C., Zhu, R., Liu, K., Su, L., Bai, B., Zhang, X., 2012. Tight gas sandstone
reservoirs in China: characteristics and recognition criteria. Journal of Petroleum
Science and Engineering 88-89, 82-91.

22
ACCEPTED MANUSCRIPT

Figure captions
Figure 1. (A) Location map of the study area and sub-tectonic units of the Songliao Basin
() Western Slope Zone, () Northern Pitching Zone, () Central Depression Zone, ()
Northeastern Uplift Zone, () Southeastern Uplift Zone, () Southwestern Uplift Zone, ()
Kailu Depression Zone; (B) The sub-tectonic units of the study area and well locations.
(From Xi et al., 2015).
Figure 2. Generalized Mesozoic-Quaternary stratigraphy of the Songliao Basin, showing
major oil and gas combinations (Form Xi et al., 2015).
Figure 3. Rock composition, lithology and sorting distribution of the K 1q4 tight sandstone
reservoirs: A, Classification of sandstone using Folks (1974) classification; B, lithology
distribution; C, sorting distribution.
Figure 4. Characteristics of the K1q4 sandstone reservoir properties: A, Porosity
distribution; B, Permeability distribution; C, Typical capillary pressure curve; D, Pore-throat
radius distribution.
Figure 5. Pore types and their vertical distribution characteristics. FD-feldspar dissolution.
Figure 6. Quartz cement characteristics in K 1q4 tight sandstone reservoirs: A, Micrograph of
thin section showing quartz overgrowth; B, Micrograph of thin section showing pore-
filling authigenic quartz; C, Idem with B but CL micrograph showing the detrital quartz
(brightly luminescent) and quartz cement (darkly non-luminescent) and microscopic stylolites;
D, Ordered mixed-layer illite/ smectite and authigenic quartz; E, Kaolinite, ordered
mixed-layer illite/ smectite and authigenic quartz; F, Illite, chlorite and authigenic quartz. QA-
authigenic quartz; QD-detrital quartz grains; I/S-mixed-layer illite/ smectite; I-illite;
K-kaolinite; Ch-chlorite; Ms- microscopic stylolites.
Figure 7. The vertical distribution characteristics of quartz cement, carbonate cement,
feldspar dissolution and intergranular volume.
Figure 8. The characteristics of feldspar dissolution and carbonate cementation: A,
Micrograph of thin section showing the feldspar partly and completely dissolved; B,
Micrograph of SEM showing feldspar dissolution, authigenic kaolinite and quartz cement; C,
Micrograph of SEM showing feldspar dissolution and authigenic albite; D, Micrograph of thin
section showing the ferrocalcite cementation; E, Micrograph of thin section showing the
ankerite cementation; F, CL micrograph of carbonate cements; G, Micrograph of thin section
showing ferrocalcite cements around the euhedral quartz overgrowth and partly replacing the
authigenic quartz; H, Micrograph of thin section showing carbonate cement replacing quartz
overgrowth; I, Micrograph of thin section showing ferrocalcite filling feldspar dissolution
pores.
QD-detrital quartz; QA-authigenic quartz; FD-feldspar dissolution; K-kaolinite; AL-albite;
Fc-ferrocalcite; An-ankerite.
Figure 9. Stratigraphy and reservoir quality of Well Qian 223 and Well Rang 24. It shows
that the carbonate cements develop more intensely along the sandstone and mudstone contact
23
ACCEPTED MANUSCRIPT

surface.
Figure 10. Figure 10. The characteristics of minor cements and clay minerals: A,
Micrograph of SEM showing pyrite framboids filling the intergranular pore; B, Micrograph of
thin section showing feldspar overgrowth and ankerite replacement; C, Micrograph of SEM
showing authigenic quartz, kaolinite and some illite; D, Micrograph of SEM showing mixed-
layer illite/smectite; E, Micrograph of SEM showing fibrous illite; F, Micrograph of SEM
showing rosette-shaped chlorite.
Py-pyrite; FA-feldspar overgrowth; An-ankerite; I/S- mixed-layer illite/smectite; I-illite;
QA-authigenic quartz; Ch-chlorite.
Figure 11. The vertical distribution characteristics of clay minerals.
Figure 12. Oil presence forms in K1q4 sandstone reservoirs from thin section and
fluorescence observation.
Figure 13. Histograms of homogenization temperature (T h) for aqueous inclusions in quart
overgrowth and pores filling authigenic quartz in K1q4 tight sandstone reservoirs.
Figure 14. The distribution characteristics of isotopes: A, Introduction of carbonate and
oxygen isotope distribution (modified from Irwin et al., 1977; Guo and Wang, 1999); B, 13C
values increase with the increasing distance of samples to source rocks.
Figure 15. Comparison of the homogenization temperatures of the aqueous inclusions in
quartz cements, carbonate cements and hydrocarbon inclusions.
Figure 16. Burial, thermal, diagenetic history and average thin section porosity evolution
trend of the K1q4 sandstone reservoirs.
The width of the symbols represents the relative occurring rate of authigenic minerals and
oil-charging.
Figure 17. Relationships between detrital grains and intergranular volumes in K 1q4
sandstone reservoirs.
Figure 18. Plot of intergranular volume (IGV) versus volume of cement in K 1q4 sandstone
reservoirs (modified from Houseknecht, 1987 and Lundegard, 1992).
Figure 19. Relationships between reservoir quality and the distance from the contact
surface of sandstone and mudstone.
Figure 20. Relationship between reservoir properties and different clay minerals.
Figure 21. Relationship between oil saturation, carbonate content and the relative content
of chlorite in Well Rang 59.

Figures

24
ACCEPTED MANUSCRIPT

Figure 1.

25
ACCEPTED MANUSCRIPT

Figure 2.

26
ACCEPTED MANUSCRIPT

Figure 3.

Figure 4.

27
ACCEPTED MANUSCRIPT

Figure 5.

Figure 6.

28
ACCEPTED MANUSCRIPT

Figure 7.

Figure 8.

29
ACCEPTED MANUSCRIPT

Figure 9.

Figure 10.

30
ACCEPTED MANUSCRIPT

Figure 11.

Figure 12.

31
ACCEPTED MANUSCRIPT

Figure 13.

Figure 14.

Figure 15.

32
ACCEPTED MANUSCRIPT

Figure 16.

33
ACCEPTED MANUSCRIPT

Figure 17.

Figure 18.

34
ACCEPTED MANUSCRIPT

Figure 19.

35
ACCEPTED MANUSCRIPT

Figure 20.

Figure 21.

36
ACCEPTED MANUSCRIPT

Table captions
Table 1. Mineralogical and isotopic composition of carbonate cements, and calculated
formation temperature of cements in the K 1q4 sandstone reservoirs. An-ankerite; Fc-
ferrocalcite.
Table 2. Microthermometric data of the aqueous fluid inclusions in K 1q4 sandstone
reservoirs.
Th-homogenization temperature; Tm-final ice melting temperature.

Tables
Table 1.
Temperatur Temperatur

Dep Carbo
W Carbonate 13CPD 18OPD Z e, W Dep 13CPD 18OPD Z e,
th, nate
ell minerals B, B, value 18OSMOW= ell th, m B, B, value 18OSMOW=
m minerals
7 7

C 228 10 R 186 100% 10

100% An -3.6 -20.6 120.73 -9 -18.2 83.79


10 0.75 9.67 24 9.3 Fc 9.67

C 228 10 R 187 100% 10

100% An -3.7 -20.7 121.75 -9.9 -21.1 110.75


10 4.49 9.41 24 2.2 Fc 9.41

C 230 10 R 187 100% 10

100% An -4.7 -20.6 120.73 -6.3 -19.5 95.13


10 2 7.42 24 2.4 Fc 7.42

C 231 10 R 187 100% 10

100% An -6 -20.9 123.83 -9.7 -22.7 128.56


10 1.15 4.60 24 2.7 Fc 4.60

C 231 10 R 187 100% 10

100% An -5.6 -20.6 120.73 -10.2 -21.4 113.90


10 2.26 5.57 24 4.02 Fc 5.57

C 231 10 R 187 100% 10

100% An -5.9 -20.9 123.83 -6.5 -18.2 83.78


10 2.37 4.81 24 5.12 Fc 4.81

C 231 10 R 187 100% 10

100% An -5.1 -18.2 98.35 -9.4 -18.4 85.46


10 6.04 7.79 24 5.63 Fc 7.79

C 228 92. R 187 100% 92.

100% Fc -11.9 -20.2 101.72 -11.4 -22.1 121.59


19 2.45 87 24 8.35 Fc 87

C 228 60% 97. R 204 100% 97.

-9.4 -21.6 116.05 -3.3 -21.8 118.24


19 4.35 Fc+40% An 29 59 0.13 Fc 29

C 228 99. R 204 100% 99.

100% Fc -8.2 -21.5 114.97 -3.7 -19 90.64


19 5.15 80 59 4.74 Fc 80

C 210 10 R 204 100% 11

100% Fc -5.8 -19.5 95.13 -2.9 -18.3 84.61


21 7.4 5.71 59 7.47 Fc 2.25

C 212 98. R 205 100% 10

100% Fc -9.1 -20.5 104.66 -6.2 -21.6 116.05


21 7.3 45 59 4.49 Fc 3.85

37
ACCEPTED MANUSCRIPT

C 213 98. R 209 100% 95.

100% Fc -9.2 -20.7 106.66 -10.5 -21.4 113.90


21 2.1 15 59 9.93 Fc 14

C 214 10 R 210 100% 95.

100% An -5.8 -20.4 118.71 -10.7 -19.9 98.86


21 4 5.26 59 0.62 Fc 48

C 214 11 R 210 100% 92.

100% An -1.1 -20.2 116.72 -11.6 -22.2 122.73


21 5.8 4.99 59 1.19 Fc 49

C 214 98. R 210 100% 94.

100% An -9 -20.5 119.71 -11 -21.6 116.05


21 9.25 66 59 1.76 Fc 02

G 119 11 R 210 100% 93.

100% Fc -2.2 -20.5 104.66 -11 -21.9 119.35


27 5.64 2.59 59 3.11 Fc 87

G 120 11 R 210 100% 95.

100% Fc -0.9 -19.4 94.22 -10.2 -21.6 116.05


27 4.6 5.80 59 7.27 Fc 65

G 120 11 R 210 100% 92.

100% Fc -1.3 -19.9 98.56 -11.5 -21.9 119.35


27 8.42 4.73 59 7.37 Fc 84

G 122 11 R 210 100% 95.

100% Fc -2.6 -20 99.80 -10.4 -20.8 107.67


27 6.4 2.02 59 8.2 Fc 64

G 123 11 R 210 100% 96.

100% Fc -1.3 -20 99.80 -10 -20.6 105.65


27 2.16 4.68 59 9.23 Fc 56

G 123 11 R 211 100% 96.

100% Fc -1.2 -19.9 98.56 -10.1 -20.9 108.68


27 3.62 4.93 59 0 Fc 21

G 150 10 R 211 100% 95.

100% Fc -5.1 -19.4 94.22 -10.1 -21.8 118.24


31 3.45 7.19 59 1.14 Fc 76

G 150 10 R 211 100% 97.

100% Fc -7.2 -18.7 88.01 -9.4 -21 109.71


31 6.4 3.24 59 1.87 Fc 59

G 151 90% 10 R 211 100% 93.

-7.8 -20.2 102.65 -11.1 -22.3 123.87


31 5.5 Fc+10% An 1.27 59 2.52 Fc 46

G 151 10 R 211 100% 94.

100% Fc -7.4 -18.4 85.46 -10.6 -21.9 119.35


31 8.1 2.98 59 2.69 Fc 69

G 151 10 R 211 100% 99.

100% Fc -6.1 -18.6 87.16 -9.3 -18.4 85.46


31 9 5.54 59 6.41 Fc 09

G 152 85% 10 R 211 100% 95.

-6.6 -19.4 95.33 -10.7 -20.6 105.65


31 0.2 Fc+15% An 4.12 59 7.84 Fc 13

G 152 10 R 212 100% 98.

100% Fc -6.4 -18.6 87.16 -9.2 -19.2 92.42


31 1.99 4.93 59 0.69 Fc 90

G 152 98. R 212 100% 96.

100% Fc -9 -21.7 117.14 -10 -20.2 101.72


31 4.19 06 59 1.2 Fc 76

H 252 10 R 212 100% 99.

100% Fc -4.2 -20.9 108.68 -9.1 -18.2 83.78


160 3.08 8.29 59 4.05 Fc 60

38
ACCEPTED MANUSCRIPT

H 252 10 R 212 100% 96.

100% Fc -4.4 -21.3 112.84 -10 -20.6 105.65


160 5.25 7.68 59 5.85 Fc 56

H 252 10 R 212 100% 10

100% Fc -4.5 -21.9 119.34 -8.4 -18.7 88.02


160 6.1 7.18 59 7.36 Fc 0.78

H 252 10 R 212 100% 94.

100% Fc -4.8 -22.2 122.73 -10.9 -21.9 119.35


160 6.4 6.41 59 8.59 Fc 07

H 253 94. R 213 100% 96.

100% Fc -10.6 -22.3 123.87 -9.7 -21.8 118.24


160 3.9 49 59 0.02 Fc 58

R 184 93. X 134 100% 98.

100% Fc -11.1 -22.7 128.56 -9.2 -20.5 104.66


24 6.09 26 125 0 Fc 25

R 184 97. X 134 100% 94.

100% Fc -9.3 -21.7 117.14 -10.9 -21.5 114.97


24 7.3 45 125 3 Fc 27

R 185 95. X 134 100% 94.

100% Fc -10.3 -22.2 122.73 -11 -20.9 108.68


24 3.7 15 125 3.6 Fc 36

R 185 94. X 134 100% 94.

100% Fc -10.5 -21.9 119.35 -10.6 -21.3 112.84


24 4 89 125 5.32 Fc 98

R 185 94. X 134 100% 93.

100% Fc -10.4 -22.2 122.73 -11.4 -20.3 102.69


24 4.35 95 125 6.4 Fc 84

R 186 96. X 135 100% 93.

100% Fc -10.1 -20.2 101.72 -11.3 -20.8 107.67


24 0.8 56 125 7.5 Fc 80

R 186 93. X 136 100% 93.

100% Fc -11.3 -21.8 118.24 -11.4 -20.7 106.66


24 1.3 30 125 3.57 Fc 64

R 186 91. X 136 100% 91.

100% Fc -11.9 -22.9 130.96 -12.2 -21 109.71


24 2.7 52 125 9 Fc 86

R 186 94.

100% Fc -10.7 -22.1 121.59


24 3.5 38

Note: The formula used in calculating calcite mineral temperature is 1000lncalcite-


6 2
water=2.7810 /T -2.89 (Friedman and O'Neil, 1977); the formula used in

6 2
calculating dolomite mineral temperature is 1000lndolomite-water=3.06 10 /T -3.24
(Matthews and Katz, 1977), and 1000lncarbonate-water=18O carbonate-18O water

39
ACCEPTED MANUSCRIPT

Table 2.
Depth, Size, Th, Tm, ice/ Salinity, NaCl wt.%
Well Host mineral types equiv.
m m
oC oC (from Bodnar,1993)

Rang 1938. Carbonate 310 aque 10 -2 3.39


54 1 cement ous 0

Rang 1938. Carbonate 37 aque 98 -3.6 5.86


54 1 cement ous

Rang 1938. Carbonate 25 aque 10 -3.7 6.01


54 1 cement ous 0

Rang 1938. Carbonate 56 aque 99 -3.7 6.01


54 1 cement ous

Rang 1938. Carbonate 27 aque 10 -2.1 3.55


54 1 cement ous 8

40
ACCEPTED MANUSCRIPT

Highlights:
(1) Quantitative diagenesis in the K1q4 sandstones was examined by a variety
of methods.
(2) Quartz cements formed from internal sources, and carbonate from external
sources.
(3) Diagenetic evolution history including oil emplacement process is
reconstructed.
(4) Compaction is more important than cementation in destroying the reservoir
quality.

41

Vous aimerez peut-être aussi