You are on page 1of 16

Organic Geochemistry 42 (2011) 340355

Contents lists available at ScienceDirect

Organic Geochemistry
journal homepage: www.elsevier.com/locate/orggeochem

Experimental kinetic study of organic matter maturation: Time and pressure


effects on vitrinite reectance at 400 C
Ronan Le Bayon a,, Gerhard P. Brey b, W.G. Ernst c, Rafael Ferreiro Mhlmann a
a

Institut fr Geowissenschaften, Technische Universitt Darmstadt, Schnittspahnstrasse 9, D-64287 Darmstadt, Germany


Institut fr Geowissenschaften, Abt. Petrologie und Geochemie, J.W. Goethe-Universitt, Altenhferallee 1, D-60438 Frankfurt am Main, Germany
c
Department of Geological and Environmental Sciences, Stanford University, Stanford, CA 94305-2115, USA
b

a r t i c l e

i n f o

Article history:
Received 6 July 2010
Received in revised form 26 January 2011
Accepted 28 January 2011
Available online 3 February 2011

a b s t r a c t
We carried out a laboratory rate study to elucidate and quantify the effects of time and pressure on vitrinite reectance (VR). A series of conned system maturation experiments was conducted at 400 C and
at pressures of 2, 10 and 20 kbar. Experiments were performed on dry (no water added) xylite of swamp
cypress and involved run lengths from 0 s to 25 days.
At 400 C, our experimental results demonstrate pressure and heating time to be important variables
that promote VR increase and therefore the maturation of Type III organic material. VR increases with
time at each investigated pressure. Despite rapid initial kinetics, the increase in VR decelerates with time
at each pressure. When VR < 1.44%, increasing pressure reduces the rate of VR increase and hence retards
the initial VR enhancement with time. The retarding effect of pressure on VR increase diminishes with
enhancing VR. The retardation of VR increase is insignicant for geological maturation at 400 C because
a VR of 1.44% is attained in only a few hours. When VR > 1.44%, increasing pressure counteracts the
deceleration of VR increase with time and thus greatly enhances the increase in VR with time. Evidently,
vitrinite maturation takes place rapidly in a dry conned system and does not require addition of water to
occur. The strong effect of the experimental heat-up on VR is obvious even for very short experiments and
must be corrected in kinetic analysis. The evolution of VR with heating time (t) and pressure (P) at 400 C
from an initial VR of 0% is well described by our new power law rate equation
nP;400  C

VRP; 400  C;t kP; 400  Ct

where the exponent n(P, 400 C) and the rate constant k(P, 400 C) increase with pressure. We regard this
kinetic formulation as a step toward a general equation describing VR evolution as a function of time, pressure and temperature for Type III organic matter. The potential of the power law formalism to model VR
from any starting VR and for complex metamorphic and heating time histories is shown by making explicit
directions on how to use such a kinetic equation.
2011 Elsevier Ltd. All rights reserved.

1. Introduction
Metamorphic organic petrology aims to identify, understand
and quantify processes and factors responsible for the formation
and transformation of organic geomaterials and their associated
physical and chemical properties. In particular, it is devoted to
the study of metamorphism and maturation in low to very low
temperature metamorphic terranes. There are many reasons for
this, but all are ultimately related to hydrocarbon exploration
and a fuller understanding of the Earths geodynamic evolution.
It is very difcult to carry out metamorphic studies in terranes consisting mostly of low temperature metasedimentary rocks because

Corresponding author. Tel.: +49 (0)61 51 16 65 41; fax: +49 (0)61 51 16 40 21.
E-mail address: rlebayon@geo.tu-darmstadt.de (R. Le Bayon).
0146-6380/$ - see front matter 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.orggeochem.2011.01.011

such conditions lead to sluggish reactions. However, the presence


of vitrinite phytoclasts provide an opportunity to understand
metamorphic and geodynamic evolution by determining rock
pressuretemperature (PT) conditions and the degree of maturity
with respect to hydrocarbon generation. This stems from the
recognition that vitrinite reectance (VR) measurements monitor
the maturation of organic material because VR in host rocks is
demonstrated to be correlated with coal rank (e.g., van Krevelen,
1953; Teichmller and Teichmller, 1954; Lopatin, 1971; Hood
et al., 1975; Hunt, 1979; Robert, 1980; Tissot and Welte, 1984;
Roksandic, 1986). Consequently, tremendous progress has been
made toward better understanding of vitrinite maturation and
reectance over the past 50 years. However, clarication of the
physical and chemical parameters that control vitrinite reectance
is necessary because VR provides valuable and even unique aid for
hydrocarbon exploration and insight into geodynamic evolution of

R. Le Bayon et al. / Organic Geochemistry 42 (2011) 340355

low temperature metamorphic terranes. Measurement of vitrinite


reectance using a microscope equipped with a photometer is
simple and practical. Vitrinite reectance can be measured on
phytoclasts of a few micrometers in diameter in host rocks or
experimental samples.
Several variables inuencing vitrinite reectance are already
recognized and have been quantitatively estimated by experiments. Others are poorly understood.
For decades, it has been well established that temperature is a
major factor controlling vitrinite reectance (Lopatin, 1971; Tissot
and Espitali, 1975; Waples, 1980; Tissot et al., 1987; Burnham and
Sweeney, 1989; Suzuki et al., 1993; Huang, 1996; Dalla Torre et al.,
1997; Ernst and Ferreiro Mhlmann, 2004). Because organic matter
maturation is irreversible, VR only records the maximum temperature to which the enclosing rocks were exposed. Measurement of
vitrinite reectance is therefore a fundamental tool widely used
by earth scientists and oil, gas and coal industry technologists to
estimate paleotemperature in low grade metamorphic settings
(e.g., Frey et al., 1980; Kisch, 1987; Underwood et al., 1999;
Ferreiro Mhlmann, 2001; rkai et al., 2002; Frings et al., 2004)
as well as the thermal maturity of hydrocarbon source rocks
(e.g., Durand and Espitali, 1976; Tissot et al., 1987; Johnsson
et al., 1993; Armstrong et al., 1996; Parris et al., 2003).
The roles of starting material composition, oxygen fugacity and
water content have been subject to experiments (Huang, 1996;
Ernst and Ferreiro Mhlmann, 2004). Humite impregnated with
liptinite (e.g., resinite, bituminite, exsudatinite from alginite)
shows reduced VR (Huang, 1996; Ernst and Ferreiro Mhlmann,
2004). These experimental studies conrm the retarding/suppressing effect of hydrogen rich macerals on VR reported from natural
environments by Wolf (1978), Hutton and Cook (1980), Price and
Barker (1985), Wenger and Baker (1987), Raymond and Murchison
(1991), Petersen and Rosenberg (1998) and Carr (2000). Ernst and
Ferreiro Mhlmann (2004) experimentally found that oxygen
fugacity does not affect the rate of vitrinite maturation. Other factors such as uid ow, the presence of oil and the partial pressures
of CO2, CH4 and argon were experimentally shown by Huang
(1996) to have imperceptible inuence on the kinetics of vitrinite
reectance evolution. In an effort to evaluate the role of water in
VR evolution, Huang (1996) conducted a series of lignite maturation experiments in a hydrous (sea water added) conned system,
an anhydrous (no water added) conned system and an anhydrous
system (no water added) open to argon gas. Despite a limited number of anhydrous conned experiments, this study concluded that
VR increases in conned systems at similar rates with or without
added water. In contrast, it also demonstrated that VR evolution
is retarded in an open system compared to a conned system. This
is in accord with experimental results of Landais et al. (1994),
which showed that conned and hydrous pyrolysis experiments
on Type III coal yield similar VR values. Furthermore, Monthioux
et al. (1985) concluded from their chemical studies on Type III organic matter that connement of generated gas in anhydrous
experiments produces results similar to those in hydrous experiments. This led Monthioux et al. (1985) to postulate that the presence of water is not necessary when pyrolysis is performed in a
conned system. Consequently, Huang (1996) suggested that
water and/or other pyrolysis products generated during the early
stage of organic matter pyrolysis in anhydrous conned systems
allow VR to change at rates similar to that under hydrous conned
conditions. Nevertheless, the need for a hydrous (a water-excess
system) maturation to promote VR requires clarication because
a series of dry (no water added) conned experiments to simulate
vitrinite maturation has never been carried out at xed PT
conditions on Type III organic matter. Thus, we investigate the
maturation of Type III organic matter in dry (no water added)
systems in this study. This will conrm whether the maturation

341

of Type III organic matter occurs in dry systems. This result will
be compared to a series of hydrous (water added) conned maturations in a further communication in order to identify whether
added water has an effect on VR.
While the effect of temperature on vitrinite reectance has been
well constrained, the inuence of time remains controversial. On
the one hand, it has been argued (e.g., Price, 1983; Ritter, 1984;
Barker, 1989, 1991) that vitrinite reectance evolution is stabilized
rapidly at geological timescales within only 10 years. On the
other hand, both the effective heating time and temperature have
been recognized as important (Lopatin, 1971; Hood et al., 1975;
Waples, 1980; Burnham and Sweeney, 1989; Sweeney and
Burnham, 1990; Hunt et al., 1991; Waples et al., 1992). These early
studies were supported by recent experiments (Huang, 1996; Dalla
Torre et al., 1997; Ernst and Ferreiro Mhlmann, 2004) showing
that, whereas time is important, the major variable that determines VR over geological time intervals is host rock temperature.
Based on these assumptions, several eld and experimentally calibrated models using Arrhenius kinetics were developed to predict
the evolution of vitrinite reectance (Waples, 1980; Burnham and
Sweeney, 1989; Larter, 1989; Suzuki et al., 1993; Huang, 1996;
Dalla Torre et al., 1997; Ernst and Ferreiro Mhlmann, 2004). However, improved understanding of the effect of time on vitrinite
reectance is still needed.
Although considerable efforts have been directed toward
understanding and quantifying the effect of temperature, oxygen
fugacity and starting material upon VR evolution, the effect of pressure on Type III organic matter maturation remains largely unknown. Despite the probable importance of pressure on VR
evolution, all of our information comes either from limited experiments with inconsistent results (Chandra, 1965; Hryckowian
et al., 1967; Goodarzi, 1985; Mastalerz et al., 1993; Hill et al.,
1994) or from controversial estimates from eld studies. Estimates
based on paleometamorphic conditions are poorly constrained in
vitrinite bearing terranes and their surrounding silico-calcic rocks
(e.g., Bostick, 1974; Diessel et al., 1978; Kisch, 1987; Dalla Torre
et al., 1994, 1996). Conicting observations in sedimentary basins
document both enhanced (Law et al., 1989; Hunt, 1996) and retarded (McTavish, 1978; Hao et al., 1995, 2007) VR proles with
the onset of overpressure (for a recent review, see Carr (2000)
and references therein). In general, pressure has been widely regarded as having little inuence on the maturation of Type III organic matter, at least when compared to temperature and
heating time (e.g., Teichmller and Teichmller, 1979; Stach
et al., 1982; Murchison et al., 1985; Huang, 1996; Ernst and Ferreiro Mhlmann, 2004). In summary, considerable confusion exists
about the pressure effect on vitrinite maturation. The lack of clear,
reliable information on the inuence of pressure on VR is surprising if one considers the strong pressure dependence of the well
known graphitization process from experimental studies (e.g., Bonijoly et al., 1982) and the importance of pressure as metamorphic
agent in most geological settings. Knowledge and accurate determination of the pressure effect on VR evolution kinetics is needed
to gain insight into and quantify the diverse processes that span organic metamorphic petrology. Thus, it is essential to experimentally investigate the inuence of pressure on VR evolution rates.
In addition to pressure, the effect of experimental heat-up to
reach the desired run temperature on the material during articial
maturation is still unknown. The experimental heat-up is the rapid
increase in run temperature from room temperature to the desired
nal run temperature (i.e., 400 C). To date, most of the existing
experimental studies on maturation of organic material do not
consider the possible inuences of experimental heat-up on vitrinite reectance (e.g., Huang, 1996; Ernst and Ferreiro Mhlmann,
2004). Dalla Torre et al. (1997) state that experimental heat-up
has no effect on VR as long as the nal temperature is reached in

342

R. Le Bayon et al. / Organic Geochemistry 42 (2011) 340355

<30 min. This is in contrast with the kinetic study of Huang (1996),
which predicts that VR strongly changes between 300 and 400 C
with run duration as short as 15 min. A stringent experimental
evaluation of heat-up effect on vitrinite maturation is needed.
Our main goal in this work is to elucidate and quantify the effects of time and pressure on vitrinite reectance in dry (no water
added) systems. Furthermore, this experimental study aims to assess the effects of dry maturation and experimental heat-up to the
nal temperature on VR. The research consists of a series of maturation experiments on organic matter at 400 C in a dry, conned
system at 2, 10 and 20 kbar for various run times. A kinetic analysis
of the experimental results was carried out to quantify and understand the evolution of measured vitrinite reectance as a function
of heating time and pressure. We formulated a pressure-dependent
empirical rate equation for the evolution of vitrinite reectance at
400 C and evaluated the pressure dependence of the rate law
parameters to fully assess the effect of pressure on the kinetics of
VR evolution. We provide explicit directions on how to use the
power law formalism to model VR for any starting VR and complex
metamorphic and heating time histories.
It is of special interest to formulate such a VR rate equation because it provides a step toward a general equation that describes
VR evolution as a function of time, pressure and temperature.
The advantage of this VR rate equation is that it does not require
complete understanding of the complex chemical reactions that
regulate vitrinite maturation. This general empirical kinetic equation will be elaborated in a future report. This VR rate equation will
be a useful tool to model the VRTt conditions in vitrinite bearing
sedimentary basins and to estimate the PTt conditions in vitrinite bearing metamorphic terranes occurring in various tectonic
settings (e.g., exhumed subducted terranes, collided terranes in
orogenic wedges). This will aid to gain insight into geodynamic
evolution of sedimentary and metamorphic terranes. Moreover,
this future general empirical kinetic equation will help to improve
hydrocarbon generation modeling in sedimentary basins.
2. Experimental and analytical procedures

Fig. 1. Photograph (a) and photomicrograph (b) (reected light) of the starting
material that consists of swamp cypress (Taxodioidea taxodium). The cellular
microstructure displays the cell walls consisting of telinite that surround the
inlling of the cell lumens. This xylite sample is Middle Eocene in age.

2.1. Experimental philosophy


We conducted an isothermal laboratory rate study of Type III
organic material maturation at various pressures to understand
and quantify the effects of time and pressure on vitrinite reectance. This series of maturation experiments was performed at
400 C in a conned system on dry (no water added) xylite of
swamp cypress at pressures of 2, 10 and 20 kbar. Experiments involved run lengths from 0 s to 25 days and were carried out by
employing a high pressure piston-cylinder apparatus and cold-seal
pressure vessels. We chose 400 C for our experiments to avoid
sluggish reactions leading to difculties in measuring maturation
rates. Nevertheless, experiments at lower and higher temperatures
in dry and wet (water added) systems are under way and will be
the subject of a future report.
2.2. Starting material
We used a xylite of swamp cypress (Taxodioidea taxodium)
(Fig. 1a) as starting material for the experiments because it is the
archetype of the botanical and chemical precursor of the huminite/vitrinite group macerals. The xylite sample is Middle Eocene
in age and was collected in the Helmstedt open pit mine from coal
seam 6 of the Wulfersdorf group, in the eastern Helmstedt syncline, Braunschweig mining area, northern Germany (Lenz and Riegel, 2001). The Wulfersdorf seam group represents a coastal
swamp paleoecosystem. Swamp cypresses are poorly represented

in the swamp forests and hummocks of the Wulfersdorf paleoecosystem and only rare stems and branches are found (Lenz and Riegel, 2001). The sample was stored at ambient conditions and
untreated before the experiments.
The xylite sample displays a recognizable cellular microstructure (Fig. 1b) in which cell walls consisting of telinite surround
structureless inllings of the cell lumens. The xylite sample lacks
liptinite macerals, such as resinite, sporopollenite, chlorophyllinite,
cutinite and bituminite that are known to suppress vitrinite reectance due to their high hydrogen content (for a recent review, see
Carr (2000) and references therein). Ernst and Ferreiro Mhlmann
(2004) recently conrmed vitrinite reectance suppression in
huminite by bituminite and resinite impregnation in their experimental study. In addition, the Wulfersdorf swamp cypress stems
show no oxidation and are very poor in other hydrogen rich plant
constituents, such as resin, wax and tannin. Furthermore, no algal
agglomerates occur in the sample bearing paleoecosystem. This
observation is important because algal agglomerates could suppress vitrinite reectance of the neighboring telinite due to their
high hydrogen contents (e.g., Stach et al., 1982; Carr, 2000).
We measured a low atomic H/C ratio of 1.27 0.03 in the immature xylite sample. This conrms the hydrogen poor nature of the
starting material. A vitrinite reectance VR = 0.167 0.020% was
measured in the starting material, indicating that the initial vitrinite/organic matter was extremely immature prior to the experiments. This allows the investigation of the largest possible VR

R. Le Bayon et al. / Organic Geochemistry 42 (2011) 340355

range because maturation prior to the experiment does not interfere with our results. This xylite sample of swamp cypress is an
ideal botanical, chemical and physical (low VR) starting material
on which to perform articial vitrinite maturation experiments.
The botanical origin and low H/C ratio of the sample indicate Type
III organic matter in the van Krevelen terminology. This type of organic matter is a good natural gas source, but is less productive of
oil (e.g., Stach et al., 1982; Taylor et al., 1998).
2.3. Experimental methods
The use of a high pressure piston-cylinder apparatus and coldseal pressure vessels to carry out the vitrinite maturation experiments is based on the ndings of Lewan et al. (1979), Monthioux
et al. (1985), Monthioux (1988), Horseld et al. (1989), Landais
et al. (1989), Lewan (1993) and Vandenbroucke et al. (1993) that
natural maturation of organic matter can be reproduced by laboratory experimentation only if conducted in conned systems (i.e., in
sealed gold or platinum capsules under external pressure, with or
without water). These workers found that organic matter heated
under fully conned laboratory conditions follows the same evolution path of H/C versus O/C atomic ratios as the natural coalication. In contrast, open system (under vacuum or inert gas) and
closed system (in sealed glass tubes and with water) maturation
studies do not yield geochemical data compatible with natural
systems.
2.3.1. High pressure piston-cylinder experiments
Experiments at pressures of 10 and 20 kbar with temperature of
400 C were conducted using the end loaded high pressure pistoncylinder apparatus of the Institut fr Geowissenschaften, Abt.
Petrologie und Geochemie, J.W. Goethe-Universitt, Frankfurt am
Main, Germany. We used Pt capsules and NaCl pressure-medium
assemblies (Fig. 2) in a 12.7 mm diameter piston-cylinder pressure
vessel. The capsules consisted of a cylinder and two lids manufactured from platinum tubes and sheets. The Pt lids were fabricated
by folding the rims of the Pt disks (3.6 mm diameter and 0.2 mm
thickness) at 90 by stamping the disks in a steel die. A Pt lid
was electric arc welded into one end of a Pt tube (57 mm long,
3.03 mm outer diameter and 0.2 mm wall thickness). The capsule

Pt70Rh30/Pt94Rh6 (type B)
thermocouple
Mullite thermocouple pipe

Pyrophyllite ring
Steel plug
BN ring
Crushable alumina ring
NaCl cylinder
Upper NaCl sleeve
Lower NaCl sleeve

Pressure
medium

Pt capsule
Graphite cylinder
Graphite disk

Furnace

10 mm
Fig. 2. Cross-section of the piston-cylinder assembly.

343

body was then annealed, hydraulically pressed to obtain a at bottomed capsule and cleaned prior to the introduction of the dry
starting material. The swamp cypress xylite was gently disaggregated to coarse and ne particles (0.11.0 mm in length,
0.1 mm thick) in an agate mill prior to loading into the Pt capsules. No water was added to the capsule. The xylite was lled
and pressed into the capsule almost to the rim. The depth to the
rim must be sufcient to house the second Pt lid. The rim of the
capsule was then crimped down over the rim of the inserted lid
and hydraulically pressed to cold seal the capsule rather than to
weld it. This avoids potential thermal alteration of the charge.
The cold sealed capsule (45 mm long after closure) was placed
vertically in a NaCl assembly acting as sample holder and pressure
medium. NaCl was chosen as pressure medium because it has good
isotropic mechanical behavior at the experimental temperature of
400 C. We machined all NaCl components by hydraulically pressing dried NaCl of 99.99% purity at a maximum of 8 kbar in steel
dies to reach at least 95% of the density of crystalline NaCl. Each
assembly consisted of a NaCl outer cylinder, a graphite heater
and an upper and a lower inner NaCl sleeve (Fig. 2). The lower inner NaCl sleeve housed the run capsule and the upper sleeve
housed the thermocouple. A cylindrical graphite resistance furnace
was used as heater. A graphite disk was added at the bottom of the
lower NaCl sleeve and graphite cylinder to improve formation of a
reliable heating circuit through the cell. The sample assembly was
placed in the tungsten carbide core of the pressure cell. The entire
assembly and thermocouple plate were then added to the pistoncylinder apparatus. To control run temperature, a Pt70Rh30Pt94Rh6
(type-B) thermocouple was inserted axially into the assembly and
located at a distance of 0.6 0.03 mm above the at top of the Pt
capsule. Thermocouple wires were placed inside a two-bore mullite pipe to isolate them from electric contact with the thermocouple plate and sample assembly. Experiments were pressurized to
the desired run pressure at room temperature and then heated to
400 C at a rate of 50 C/min. This led to a heat-up time of
<8 min. Temperature ramp rate and temperature during the experiment were controlled automatically via a programmable EUROTHERM power controller. Temperature was held constant to
within 1 C during experiments. Temperatures reported in this
experimental study are uncorrected for potential thermal gradients in the capsule. Stated values are thought to be accurate to
1 C. Pressure was kept constant for the duration of the experiment by manual adjustment if necessary. Pressure is considered
to be accurate to 1 kbar. Run conditions were maintained for 0 s
to 25 days. Experiments were quenched in <10 s by switching off
the power to the heater whilst maintaining quasi run pressure.
Afterwards, pressure was decreased in 30 min.
2.3.2. Cold-seal pressure vessel experiments
Several unsuccessful attempts were made to perform experiments at 2 kbar and 400 C using the high pressure piston-cylinder
apparatus: The run capsules exploded. To overcome this problem,
an alternative experimental method was used for low pressure
experiments. Experiments at a xed pressure of 2 kbar were conducted in cold-seal pressure vessels in the hydrothermal laboratory of the Geological and Environmental Sciences Department of
Stanford University, California. The xylite was loaded into gold capsules. Run capsules were placed in cone-in-cone cold-seal pressure
vessels (Kerrick, 1987) connected to a water pressure reservoir, a
Heise bourdon tube gauge and an air driven hydraulic pump. Prior
to beginning this study, each autoclave and thermocouple assembly was temperature calibrated against the melting point of NaCl
at 1 atm. After pressurization of an experimental sample to 2 kbar
aqueous uid pressure, the vessel was placed in a nichrome wound
resistance furnace and externally heated to the desired temperature using an electronic proportional temperature regulator

344

R. Le Bayon et al. / Organic Geochemistry 42 (2011) 340355

connected to a chromelalumel sensing thermocouple. The sample


temperature was read from the EMF a run thermocouple inserted
in an external well in the hot end of the autoclave. Heating and
cooling times for experiments were 30 min. Pressure and temperature uncertainties were 25 bar and 2 C in cold-seal pressure
vessels. Run conditions were maintained for 0 s to 25 days.
2.4. Analytical methods
After the experiments, the capsules were recovered and cut diametrically in half for vitrinite reectance analysis. The preparation
of vitrinite mounts consisted of plucking numerous vitrinite particles from the recovered samples, embedding them in epoxy resin
and then polishing the mounts at room temperature using 3, 1
and 0.5 lm diamond powder and felted wool for the last polish
to avoid scratches that cause relief and shadows that change reectance. The same procedure of preparation for vitrinite mounts was
applied to the initial material prior to experiments. VR measurements were performed at the Institut fr Geowissenschaften of
the Technische Universitt Darmstadt, Germany. Vitrinite reectance was measured according to standard procedures (Stach
et al., 1982). We measured the 546 nm wavelength monochromatic light reected from the vitrinite mount surface using a photomultiplier coupled to a Leitz Orthoplan-photometer microscope
calibrated with a single standard. The microscope was equipped
with a 125 oil immersion objective coupled with a 10 ocular.
A non-drying immersion oil with a refraction index ne = 1.518 (at
23 C and for 546 nm wavelength monochromatic light) was used
for reectance measurements. The VR measurement equipment
was placed on a sturdy work bench to avoid vibration during
measurement. In addition, measurements were operated in dim
light at about 21 C. Calibration was done according to standard
procedures detailed by Taylor et al. (1998). Standards used for calibration (with corresponding reectance R) were: Yttriumaluminiumgarnet (R = 0.880%), gadoliniumgalliumgarnet (R = 1.719%)
and cubic zirconia (R = 3.114%). The quality of calibration was
checked every two hours during measurements. To calculate the
VR mean value of an experimental sample, we conducted a
minimum of 100 VR measurements. For vitrinite reectance below
1.3%, measured VR is the mean random vitrinite reectance. For
vitrinite reectance above 1.3%, measured VR is the maximum
vitrinite reectance determined under polarized light. Vitrinite
reectance measurements were only carried out on telinite.
3. Experimental results
3.1. Microscopic observations
Prior to discussion of the VR results, it is necessary to report a
few observations concerning microscopic evolution of the xylite
matured at 400 C as a function of pressure and time to expand
our limited knowledge of isothermal maturation of Type III organic
matter. Before the experiments, the initial xylite displayed the easily recognizable cellular microstructure of swamp cypress: Bright
(on a grey scale) cell walls consisting of telinite surrounding the
dark inlling of the cell lumens (Fig. 1b). After the experiments,
several critical aspects of the organic maturation were observed
in the experimentally matured xylite.
In all 2 kbar experiments with heating times P1 h, we found
inhomogeneous strong granular disintegration of the material to
a ne mosaic structure typical of coking (Fig. 3h). This decomposed
material was difcult to measure because the grains were smaller
than the aperture of the photo-multiplier. Similar observations
document such a coking effect at 2 kbar and 400 C (Ernst and
Ferreiro Mhlmann, 2004). Nevertheless, some rare matured xylite

Fig. 3. Photomicrographs (reected light) showing the textural relations in run


products from experiments at 400 C and at different effective heating time and
pressure conditions. (a) 0 s20 kbar (run RLB-e84), (b) 15 min20 kbar (run RLBe26), (c) 1 day20 kbar (run RLB-e07), (d) 25 days20 kbar (run RLB-e01), (e)
15 min10 kbar (run RLB-e22), (f) 1 day10 kbar (run RLB-e27), (g) 25 days
10 kbar (run RLB-e44), (h) 1 h02 kbar (run GE-V8). T = telinite (cell wall); CL = cell
lumen; GD = granular decomposition of the organic material.

particles or parts of them isolated in the coke were preserved from


granular disintegration. Some initial cellular (telinite and cell inlling) microstructures of the cypress were recognized in these preserved grains, where VR was carefully measured. In the 2 kbar
runs with heating times <1 h, no disintegration/coking occurred
and the original botanical structure was preserved.
In all 10 and 20 kbar experiments, the initial cellular microstructure of the wood (cell walls and cell lumens) was recognizable
(Fig. 3ag). There was no general gelication and therefore the cell
walls consisting of telinite were readily distinguished from the

345

R. Le Bayon et al. / Organic Geochemistry 42 (2011) 340355

inlling of the cell lumens. This distinction was straightforward in


all experiments at 10 and 20 kbar and for durations 61 day. The
only deformation of the material occurred during the lling of
the experimental capsule prior to high pressure experiments (folding of the wood tissue). Telinite and cell inlling brightened with
run duration during isobaric experiments at 10 and 20 kbar. The
optical distinction between telinite and cell inlling became more
difcult with increasing run time. Hence, more attention was
needed to distinguish telinite from the inlling of the cell lumens
for experiments with duration >1 day at 10 (Fig. 3g) and 20 kbar
(Fig. 3d). However, the original cellular microstructure of the cypress was still recognizable after 25 days of high pressure experimental maturation.
3.2. Vitrinite reectance
VR results from dry (no water added) experiments at 400 C and
at three different pressures (2, 10 and 20 kbar) are presented in
Table 1 and plotted in VR versus effective heating time t0 diagrams
(Fig. 4). Besides our experimental results, we use published VR data
obtained at 2 kbar and 400 C from a previous experimental study
by Ernst and Ferreiro Mhlmann (2004) who experimented on the
same kind of starting material. They used humite from an angiosperm lignite as starting material.
VR results from duplicated runs carried out in this experimental
program at various Pt0 conditions (Table 1 and Fig. 4) are consistent in most cases within the limits of experimental and measurement uncertainties and validates our techniques.

3.2.1. Effect of dry experimental maturation on VR


Prior to detailed analysis of the changes in VR, we provide a general observation concerning the maturation of Type III organic matter in a dry (no water added) system. The rst impression from
results presented in Table 1 and Fig. 4ac is that VR increases with
run time and with pressure at 400 C. This observation is of prime
importance because it demonstrates that Type III organic matter
maturation occurs in the dry system and addition of water is not
required at 400 C. A similar observation has been reported by
Monthioux et al. (1985), Landais et al. (1994) and Huang (1996).
3.2.2. Effect of experimental heat-up to T = 400 C on VR
To date, no attention has been paid to the effect of experimental
heat-up procedures on organic matter maturation to the nal run
temperature and pressure. This is surprising in the light of the sensitivity of organic matter maturation to temperature. For such an
evaluation, we performed experiments at 2, 10 and 20 kbar, raised
the temperature to 400 C and quenched immediately. We measured a signicant increase in VR in all three run products (Table 1
and Fig. 4ac). From the immature starting material with
VR = 0.17%, VR changes during the 8 min of heat-up time to
0.504% at 20 kbar (RLB-e84) and 0.67% at 10 kbar (RLB-e43). At
2 kbar and after 30 min of heat-up time VR = 0.745% (GE-V4). It is
to be noted that, the vitrinite run products surpass the onset of
oil generation commonly xed at VR = 0.5% (e.g., Taylor et al.,
1998) after these very short heat-up times. These results demonstrate the strong effect of the experimental heat-up on vitrinite
reectance and therefore on organic matter maturation, even with

Table 1
Experimental maturation conditions and results.a
Run no.

Methodb

P (bar)

t0 (s)

VR(P, 400 C, t0 ) (%)

rVR(P, 400 C, t )

Source

GE-V4
GE-V3
GE-V8
GE-V12
GE-H28
GE-F30
GE-H30
GE-F27
GE-H27
RLB-e43
RLB-e30
RLB-e39
RLB-e22
RLB-e33
RLB-e16
RLB-e42
RLB-e41
RLB-e40
RLB-e27
RLB-e05
RLB-e19
RLB-e44
RLB-e84
RLB-e35
RLB-e26
RLB-e02
RLB-e06
RLB-e03
RLB-e08
RLB-e07
RLB-e04
RLB-e36
RLB-e01

CSPV
CSPV
CSPV
CSPV
CSPV
CSPV
CSPV
CSPV
CSPV
HPPC
HPPC
HPPC
HPPC
HPPC
HPPC
HPPC
HPPC
HPPC
HPPC
HPPC
HPPC
HPPC
HPPC
HPPC
HPPC
HPPC
HPPC
HPPC
HPPC
HPPC
HPPC
HPPC
HPPC

2000
2000
2000
2000
2000
2000
2000
2000
2000
10,000
10,000
10,000
10,000
10,000
10,000
10,000
10,000
10,000
10,000
10,000
10,000
10,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000

0
900
3600
86,400
1,555,200
2,505,600
2,505,600
4,320,000
4,320,000
0
100
240
900
900
3600
3600
8100
18,000
86,400
432,000
432,000
2,160,000
0
900
900
3600
3600
18,000
18,000
86,400
432,000
432,000
2,160,000

0.745
0.753
1.176
1.624
2.041
2.086
2.068
2.116
2.191
0.669
0.666
0.683
0.864
0.862
1.066
0.952
1.262
1.386
1.695
2.098
2.017
2.561
0.504
0.681
0.706
0.860
0.883
1.129
1.130
1.879
2.326
2.130
3.360

0.044
0.039
0.042
0.048
0.057
0.094
0.058
0.072
0.075
0.034
0.037
0.047
0.053
0.038
0.041
0.054
0.078
0.031
0.060
0.037
0.031
0.057
0.027
0.033
0.023
0.024
0.036
0.039
0.048
0.034
0.018
0.053
0.051

137
442
100
101
100
34
100
41
100
239
173
116
278
100
500
500
223
263
100
167
161
235
500
136
115
100
100
137
500
100
200
100
105

c
c
c
d
d
d
d
d
c
c
c
c
c
c
c
c
c
c
c
c
c
c
c
c
c
c
c
c
c
c
c
c

a
Notations: P, pressure; t0 , experimental effective heating time; VR(P, 400 C, t0 ), measured vitrinite reectance after an experimental effective heating time t0 at 400 C and
pressure P; rVR(P, 400 C, t0 ), standard deviation of VR(P, 400C, t0 ); N, number of measurements.
b
HPPC, high pressure piston-cylinder apparatus experiments; CSPV, cold-seal pressure vessel experiments.
c
Results of this study.
d
Results of Ernst and Ferreiro Mhlmann (2004).

346

R. Le Bayon et al. / Organic Geochemistry 42 (2011) 340355

3.0

VR (10 kbar, 400 C, t' ) (%)

VR (2 kbar,400 C, t' ) (%)

2.5

2.0

1.5

2 kbar
1.0

0.5
: VR heat-up (2 kbar,400 C)

r 2 = 0.997

0.0

2.5
2.0
1.5

10 kbar

1.0
0.5
: VR heat-up (10 kbar,400 C)

r 2 = 0.996

0.0
0

10

20

30

40

50

60

10

t' (days)

2.5
2.0

20 kbar

1.5
1.0
r2

1.2

3.0

0.5

20

25

30

1.3

c
VR (P, 400 C, t' ) (%)

VR (20 kbar,400 C, t' ) (%)

4.0
3.5

15

t' (days)

: VR

= 0.995

heat-up

: VR heat-up (10 kbar,400 C)


: VR heat-up (2 kbar,400 C)

1.1
1.0
0.9
0.8
: 10 kbar

0.7

: 2 kbar

(20 kbar,400 C)

0.6

0.0
0

10

15

20

25

30

t' (days)

1000

2000

3000

4000

t' (s)

Fig. 4. Vitrinite reectance VR as a function of experimental effective heating time t0 at 400 C and (a) 2 kbar, (b) 10 kbar and (c) 20 kbar. The curves are the least-squares best
0
ts to the data using the power law equation VR(P, 400 C, t0 ) = VR(P, 400 C, 0) + (k0 (P, 400 C) t0 )n (P, 400 C) (Eq. (1)). The heat-up vitrinite reectance VRheat-up(P, 400 C) (star) is
shown at each investigated pressure for information but is excluded from this kinetic t. VR data similar to their respective starting vitrinite reectance VRheat-up(P, 400 C) are
neither plotted nor considered into this kinetic t (see text). The error bars are one standard deviation rVR(P, 400 C, t0 ). (d) VR(P, 400 C, t0 ) of the shorter experimental effective
heating times t0 as a function of t0 at 400 C and 2 and 10 kbar. VR(P, 400 C, t0 ) plateaus that indicate short activation time before VR increases from the starting vitrinite
reectance VRheat-up(P, 400 C) are clearly visible.

very short duration. This increase in VR during heat-up to 400 C is


described as VRheat-up(P, 400 C) below. This heat-up vitrinite reectance VRheat-up(P, 400 C) must be taken as the starting point in our
kinetic study and not the initial VR = 0.17% of the starting material.
It is also important to note that VRheat-up(P, 400 C) decreases with
increasing pressure.
3.2.3. Qualitative effects of effective heating time on VR
The results presented in Table 1 and Fig. 4ac (diagrams of VR
versus t0 ) show an isobaric increase in VR with effective heating
time at 400 C and at each investigated pressure. The isobaric rate
of VR increase is initially very rapid as seen from the steep positive
slopes of the VR versus t0 relation in Fig. 4ac. Nevertheless, the reduced slopes of the VRt0 versus with t0 trend (Fig. 4ac) indicate
slowing of the rate of increase in VR with run duration at each pressure and thus a deceleration of VR enhancement with t0 . These
observations conrm previous experimental results (Huang,
1996; Ernst and Ferreiro Mhlmann, 2004).
3.2.4. Qualitative effects of pressure on VR
As seen in Fig. 4ac, pressure has a strong effect on the kinetics
of VR evolution. The isobaricisothermal VR evolution trends with
t0 show that the increase in VR with t0 is greatly enhanced with
pressure. This observation is supported by the lesser reduction in
VRt0 slope with increasing pressure. It illustrates that increasing
pressure leads to a lesser deceleration of VR enhancement with

time. These important qualitative observations on the effects of


pressure on VR evolution rates are further conrmed by the decrease in VRheat-up(P, 400 C) with pressure (see above). Consequently, elevated pressure counteracts the decrease in the rate of
VR increase with time. Evidently, pressure promotes VR increase
for Type III organic material, conrming our preliminary results
(Le Bayon et al., 2007). The rate of vitrinite reectance increase is
not only a thermally activated process, it is also pressure
controlled.
3.2.5. Short activation time
The results of experiments RLB-e30 and RLB-e39 at 10 kbar and
GE-V3 at 2 kbar (Table 1 and Fig. 4d) provide evidence for a short
activation time before measurable maturation starts at 400 C with
a VRheat-up(P, 400 C) starting material. No experimental activation
time was detected at 20 kbar. The existence of such an activation
time is demonstrated at 2 and 10 kbar by the plateaus of vitrinite
reectance data similar to their respective starting vitrinite reectance VRheat-up(P, 400 C) with increasing run duration at 400 C.
These plateaus indicate that at least 15 min at 2 kbar and 4 min
at 10 kbar are required before a change in VR of the run material
is detected. However, it is impossible to deduce any trend in the
activation time evolution with pressure because the maturation
experiments are conducted with a starting material having different VRheat-up(P, 400 C) at each experimental pressure. In addition,
it is experimentally difcult to retrieve information on any

347

R. Le Bayon et al. / Organic Geochemistry 42 (2011) 340355

variation of the activation time with VRheat-up(P, 400 C) evolution


of the starting material. Thus, the concept of activation time is
complex. It is judicious to ignore any systematic treatment of the
activation time as a function of pressure or VRheat-up(P, 400 C) in
our kinetic analysis. Nevertheless, the VRheat-up(P, 400 C) and activation time effects on VR evolution with heating time and pressure
have to be corrected in our following kinetic analysis. Furthermore,
the VRheat-up(P, 400 C) data and the VR data showing the short activation time before to detect an increase in maturation with run
duration at each investigated pressure and 400 C are excluded
from the following kinetic analysis. This stems from the lack of precise time constraints for the start of increase in maturation of a
VRheat-up(P, 400 C) starting material with run duration at each
pressure and 400 C.

These curves show the expected larger increase in VR with time


and the lesser deceleration with time of VR isothermal increase with
increasing pressure that were described above. In addition, these
curves display the very short activation times that were noted
above (Fig. 4d).
Despite the excellent correlation coefcients obtained by using
Eq. (1) to t the experimental results, caution is required to extrapolate this rate equation to nature or to compare the kinetic parameters in Table 2. Indeed, we have pointed out above that the
vitrinite maturation rate experiments at 400 C and xed pressure
were conducted using a starting material having vitrinite reectance VRheat-up(P, 400 C). This value was gained during the experimental heat-up to 400 C at pressure P. Therefore, Eq. (1)
describes the evolution of VR with time and pressure at 400 C
for a starting material with vitrinite reectance VRheat-up(P,
400 C). In addition, the 2 and 10 kbar runs have an activation
time before detecting maturation. This activation time effect is
dened for a VRheat-up(P, 400 C) and a pressure P. Therefore, the
VRheat-up(P, 400 C) and activation time effects restrict the applicability of Eq. (1) for geological settings. Consequently, we must correct Eq. (1) for the VRheat-up(P, 400 C) and activation time effects.

4. Kinetic analysis
4.1. Empirical rate equation for VR evolution
In order to quantify and understand the relationship between
vitrinite reectance, pressure and time, we performed a kinetic
analysis of the experimental results. This analysis will be a useful
tool for extrapolating vitrinite reectance to geological situations
via an empirical rate equation.

4.1.2. Corrections to the basic rate equation


To overcome the VRheat-up(P, 400 C) and activation time effects,
our strategy is based on an analytical correction of Eq. (1) resulting
from two considerations. First, it is experimentally impossible to
get vitrinite reectance below VRheat-up(P, 400 C) because this VR
was obtained for the experimental effective heating time t0 = 0 s.
Second, we have no information about the activation time evolution with pressure and with the same starting material after
heat-up to 400 C.
We dene a new base of effective heating time, t, which permits
quantication of VR evolution versus corrected effective heating
time t with a starting VR equal to 0% at t = 0, i.e. VR(P, 400 C,
t = 0) = 0%. This is done by adding or subtracting time to the t0 base
at each pressure so that a power law equation similar to Eq. (1) has
its original VR at t = 0 (VR(P, 400 C, t = 0)) equal to 0%. We arbitrarily dene the starting VR equal to 0% at t = 0 despite the fact
that the vitrinite precursor has a reectance ranging between
0.1 and 0.2%. The choice to t our experimental data and to formulate a VR evolution rate equation with a starting VR equal to 0%
allows a practical VR calculation from any value of starting VR with
our VR evolution rate equation once a simple correction of the
heating time is performed (see Section 4.4).
Finally, corrections for the VRheat-up(P, 400 C) and activation
time effects require the addition of 15 s to t0 at 20 kbar and subtraction of 430 s from t0 at 10 kbar and 2780 s from t0 at 2 kbar
in order to transform the effective heating time t0 base to the corrected effective heating time t base. We list in Table 3 the corrected
effective heating time t for the VR data considered in the kinetic
analysis, i.e., VR data showing an increase in maturation from
VRheat-up(P, 400 C) (see above).

4.1.1. The basic rate equation


We obtained experimental VR data over a sufciently wide
range of time to statistically discriminate a rate law that describes
VR evolution with effective heating time. The experimental VR data
employed for the kinetic analysis are given in Table 1. Nevertheless, the starting vitrinite reectance data VRheat-up(P, 400 C) and
the VR data showing the short activation time before to detect an
increase in maturation with run duration at each investigated pressure and 400 C were excluded from the following kinetic analysis
(see above). Consequently, the VR data RLB-e84 at 20 kbar, RLBe43, RLB-e30 and RLB-e39 at 10 kbar and GE-V4 and GE-V3 at
2 kbar are not considered in the kinetic analysis. We best t our
experimental VR data over effective heating time at each run pressure P and 400 C by least squares to an empirical equation of the
form
0

VRP; 400  C;t0 VRP; 400  C;0 k P; 400  Ct 0 n P;400

 C

where VR(P, 400 C, t ) is the vitrinite reectance value after effective heating time t0 (s), VR(P, 400 C, 0) is the original vitrinite reectance at t0 = 0, k0 (P, 400 C) is the rate constant with dimensions
[time]1 and n0 (P, 400 C) is the dimensionless exponent at a given
pressure P and 400 C. The exponent n0 (P, 400 C) has no signicance in terms of stoichiometries of the chemical reaction mechanisms. The experimental VR data were weighted assuming the
uncertainties in measured VR given in Table 1. Eq. (1) provides a satisfactory description of our VR data as all ts yield correlation coefcients r2 > 0.995 (Table 2). In Table 2, we provide the magnitudes
of VR(P, 400 C, 0), k0 (P, 400 C) and n0 (P, 400 C) that were determined at each run pressure P by tting the experimental VR data
(Table 1) to Eq. (1). We plotted the three isobaric least squares
regression curves in VR(P, 400 C, t0 ) versus t0 diagrams (Fig. 4ac).

4.1.3. The rate equation for VR evolution


Having corrected the effective heating time for the activation
time problem and for the VRheat-up(P, 400 C) effect, we are now

Table 2
Vitrinite reectances at effective heating time t0 = 0 VR(P, 400 C, t0 = 0), rate constants k0 (P, 400 C) and power law exponents n0 (P, 400 C) at pressures P and 400 C.a
P (bar)

VR(P, 400 C, t0 = 0) (%)

rVR(P, 400 C, t =0)

k0 (P, 400 C) (s1)

rk (P, 400 C) (s1)

n0 (P, 400 C)

rn (P, 400 C)

r2

2000
10,000
20,000

1.426  10+2
6.390  101
2.788  101

2.030
6.078  101
6.250  102

6.675  10+160
6.678  1023
6.558  1013

1.790  10+308
5.648  1022
5.014  1013

9.586  104
9.477  102
2.584  101

3.848  105
2.694  102
1.521  102

0.997
0.996
0.995

a
Notations: rVR(P, 400 C,
correlation coefcient.

t0 =0),

standard deviation of VR(P, 400 C, t0 = 0); rk0 (P,

400 C),

standard deviation of k0 (P, 400 C); rn0 (P,

400 C),

standard deviation of n0 (P, 400 C); r2,

348

R. Le Bayon et al. / Organic Geochemistry 42 (2011) 340355

Table 3
Experimental maturation conditions and results after corrections of the effective heating time.a

Run no.

P (bar)

t (s)

VR(P, 400 C, t) (%)

rVR(P, 400 C, t)

ln t (t in s)

ln VR(P, 400 C, t)

GE-V8
GE-V12
GE-H28
GE-F30
GE-H30
GE-F27
GE-H27
RLB-e22
RLB-e33
RLB-e16
RLB-e42
RLB-e41
RLB-e40
RLB-e27
RLB-e05
RLB-e19
RLB-e44
RLB-e35
RLB-e26
RLB-e02
RLB-e06
RLB-e03
RLB-e08
RLB-e07
RLB-e04
RLB-e36
RLB-e01

2000
2000
2000
2000
2000
2000
2000
10,000
10,000
10,000
10,000
10,000
10,000
10,000
10,000
10,000
10,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000

820
83,620
1,552,420
2,502,820
2,502,820
4,317,220
4,317,220
470
470
3170
3170
7670
17,570
85,970
431,570
431,570
2,159,570
915
915
3615
3615
18,015
18,015
86,415
432,015
432,015
2,160,015

1.176
1.624
2.041
2.086
2.068
2.116
2.191
0.864
0.862
1.066
0.952
1.262
1.386
1.695
2.098
2.017
2.561
0.681
0.706
0.860
0.883
1.129
1.130
1.879
2.326
2.130
3.360

0.042
0.048
0.057
0.094
0.058
0.072
0.075
0.053
0.038
0.041
0.054
0.078
0.031
0.060
0.037
0.031
0.057
0.033
0.023
0.024
0.036
0.039
0.048
0.034
0.018
0.053
0.051

6.709
11.334
14.255
14.733
14.733
15.278
15.278
6.153
6.153
8.061
8.061
8.945
9.774
11.362
12.975
12.975
14.585
6.819
6.819
8.193
8.193
9.799
9.799
11.367
12.976
12.976
14.586

4.443
4.120
3.892
3.870
3.879
3.856
3.821
4.751
4.754
4.541
4.654
4.372
4.279
4.077
3.864
3.904
3.665
4.989
4.953
4.756
4.730
4.484
4.483
3.974
3.761
3.849
3.393

Notations: P, pressure; t, corrected effective heating time; VR(P, 400 C, t), measured vitrinite reectance at the corrected effective heating time t, 400 C and pressure P;

rVR(P, 400 C, t), standard deviation of VR(P, 400 C, t).

in a position to accommodate an initial vitrinite reectance value


in our t. Hence, we plot in Fig. 5 and best t our experimental
VR(P, 400 C, t) data without the VRheat-up(P, 400 C) data and the
VR data similar to their respective starting vitrinite reectance
VRheat-up(P, 400 C) (see above) over corrected effective heating
time (Table 3) at each run pressure P by least squares to a power
law equation of the form

VRP; 400  C;t kP; 400  CtnP;400

 C

where VR(P, 400 C, t) is the vitrinite reectance value attained after


corrected effective heating time t (s), k(P, 400 C) is the rate constant with dimensions [time]1 and n(P, 400 C) is the dimensionless exponent at a given pressure P and 400 C. This rate equation

is dened for a starting VR equal to 0% at t = 0. The experimental


VR data were weighted assuming the uncertainties in measured
VR given in Table 3. Table 4 lists the magnitudes of k(P, 400 C)
and n(P, 400 C) that were determined at each run pressure P by tting the (VR(P, 400 C, t), t) data in Table 3 to Eq. (2). We plot the
three isobaric least squares regression curves in the VR(P, 400 C,
t) versus t diagram (Fig. 5). Eq. (2) provides a satisfactory description of our VR data with time inasmuch as the ts all yield correlation coefcients r2 > 0.994 (Table 4). However, to validate our
power law Eq. (2) to t the experimental data, we transformed
the nonlinear relation between VR(P, 400 C, t) and t into a linear
relation. To effect the desired linearization, we simply take the natural logarithm of Eq. (2), yielding

ln VRP; 400  C;t nP; 400  C ln kP; 400  C nP; 400  C ln t


4.0

3
400 C

VR (P, 400 C, t ) (%)

3.5

bar

20 k

3.0

10 kbar

2.5
2.0

2 kbar

1.5
1.0
0.5

20 kbar

r 2 = 0.994

10 kbar

r 2 = 0.997

2 kbar

r 2 = 0.997

0.0
0

10

20

30

40

50

t (days)
Fig. 5. Variation in vitrinite reectance VR(P, 400 C, t) with corrected effective
heating time t at 400 C and 2, 10 and 20 kbar (Table 3). The curves are the
least-squares best ts to the data using the power law equation VR(P, 400 C, t) =
(k(P, 400 C) t) n(P, 400 C) (Eq. (2)). The error bars are one standard deviation
rVR(P, 400 C, t).

where n(P, 400 C) is the slope and n(P, 400 C) ln k(P, 400 C) is the
normalization constant of the straight line ln VR(P, 400 C, t) versus
ln t. The straight lines are calculated using Eq. (3) at 2, 10 and
20 kbar, using the best estimates for the constants n(P, 400 C)
and k(P, 400 C) given in Table 4. Results of these calculations are
presented in Fig. 6. Fig. 6 is a ln VR(P, 400 C, t) versus ln t plot that
gives a set of three lines of ln VR(P, 400 C, t) evolution with ln t at
three different pressures. The pairs (ln VR(P, 400 C, t), ln t) (Table 3)
are found to lie on or close to the straight lines shown in Fig. 6. This
graphic check demonstrates that the variables ln VR(P, 400 C, t) and
ln t are linearly related. This result conrms the use of our power
law equation (Eq. (2)) to describe the evolution of vitrinite reectance with time at different pressures and 400 C. In addition, this
linear Eq. (3) and its plot in Fig. 6 are useful because they show that
both short (65 days) and long (>5 days) duration experiments are
important to quantify VR evolution with heating time. Furthermore,
this kinetic analysis shows that VR evolution is continuous with
heating time.
We are now in a position to discuss the VR magnitude and
kinetics because Eqs. (2) and (3) and their related VRt data

349

R. Le Bayon et al. / Organic Geochemistry 42 (2011) 340355


Table 4
Rate constants k(P, 400 C) and power law exponents n(P, 400 C) at pressures P and 400 C.a
k(P, 400 C) (s1)

2000
10,000
20,000
a

rk(P, 400 C) (s1)

31

30

9.541  10
1.648  1019
2.782  1014

Notations: rk(P,

400 C),

2

3.337  10
1.523  1019
7.894  1015

standard deviation of k(P, 400 C); rn(P,

400 C),

5h

1d

5d

 nP; 400  C1 ln VRP; 400  C;t

25 d 100 d
4.0

3.0

2 kbar

2.5
2.0

4.0

1.5

ar

b
2k

1.2
ar
kb

1.0

10
ar

r 2 20 kbar = 0.994

kb

0.8

r 2 10 kbar = 0.997

20

VR (P, 400 C, t ) (%)

ln VR (P, 400 C, t)

400 C

5.0

r 2 2 kbar = 0.997

10

ln t ( t

0.997
0.997
0.994

lndVRP; 400  C;t=dt lnnP; 400  CkP; 400  C 1

10 kbar

4.5

4.460  10
3.780  103
3.101  103

standard deviation of n(P, 400 C); r2, correlation coefcient.

20 kbar
3.5

3

7.120  10
1.282  101
2.062  101

t
3
15
min min 1 h

r2

rn(P, 400 C)

n(P, 400 C)

12

14

0.6
16

in s)

Fig. 6. Double-logarithmic plot of the vitrinite reectance VR(P, 400 C, t) versus


corrected effective heating time t at 400 C and 2, 10 and 20 kbar (Table 3). The
straight lines are calculated at 2, 10 and 20 kbar, using the linear equation ln VR(P,
400 C, t) = n(P, 400 C) ln k(P, 400 C) + n(P, 400 C) ln t (Eq. (3)) and the best
estimates for the constants n(P, 400 C) and k(P, 400 C) given in Table 4.

 C1

4
Taking the natural logarithm of Eq. (4) for graphic convenience
and using k(P, 400 C)1 VR(P, 400 C, t)n(P, 400 C) in place of t from
a rearrangement of Eq. (2), we obtain by rearranging and
simplifying

2.5

4.0

10
20

kb

400 C

10

ar

10

10

10

15

20

dVR (P, 400 C, t ) / dt (%.s 1 )

t nP;400

1.0 1.44

5
r
ba
2k

 C

0.6

ar
kb

dVRP; 400  C;t=dt nP; 400  CkP; 400  CnP;400

VR (%)
0.4
10

(Table 3) are dened for a starting VR equal to 0% at t = 0. Fig. 6


points out that the highest initial VR magnitude and evolution
kinetics occur at the lowest pressure. This reveals that pressure retards the initial VR increase at 400 C. Furthermore, the increase in
VR decelerates with increasing time. However, Figs. 5 and 6 conrm our qualitative observations reported above for t > 15 h. In
particular, pressure enhances VR increase with time. This calls for
a lesser deceleration of VR increase with increasing pressure.
These observations show that the role of pressure on VR is
important and complex. It is essential to accurately analyze the effects of pressure on VR increase kinetics. For this purpose, we assess the effects of pressure on the rate at which VR increases
with time for each VR magnitude. The rate at which VR increases
with time at each pressure is found by differentiating Eq. (2) with
respect to time, yielding

We calculate the linear Eq. (5) at 2, 10 and 20 kbar using the


best estimates for the constants n(P, 400 C) and k(P, 400 C) given
in Table 4. The evolution of ln (dVR(P, 400 C, t)/dt) versus ln VR at
each pressure is presented in Fig. 7. Obviously, the rate of VR increase decreases with VR at each pressure (Fig. 7). This argues for
the deceleration with heating time of VR increase at each pressure
that is described above and shown in Figs. 5 and 6. Nevertheless,
Fig. 7 shows that increasing pressure results in the lesser deceleration of VR enhancement with VR and thus with heating time. This
leads the evolutions of the rate of VR increase with VR at 2, 10 and
20 kbar to intersect at VR = 1.44%. This calls for a variation in the
effects of pressure on the VR increase kinetics. When VR < 1.44%,
Fig. 7 demonstrates that increasing pressure diminishes the rate
of VR increase. Hence, increasing pressure results in the retardation
of the initial VR enhancement with heating time that is displayed
for t < 15 h in Fig. 6. Consequently, the lower pressure, the higher
are the initial VR magnitude and evolution kinetics. Nevertheless,
the retarding effect of pressure on VR enhancement decreases with
increasing VR and thus with heating time. This is inferred from the
decrease of the difference between the rates of VR increase at different pressure (Fig. 7). When VR > 1.44%, Fig. 7 indicates that pressure reduces the decrease in the rate of VR increase occurring with
enhancing VR. This points out that pressure counteracts the deceleration of VR increase with time. Therefore, increasing pressure
leads to the larger enhancement of VR increase with time when
VR > 1.44%. This is supported by the VRt evolutions presented in
Figs. 5 and 6 when t > 15 h. To sum up, pressure retards the

ln (dVR (P, 400 C, t ) / dt )

P (bar)

25
6

lnVR (VR in %)
Fig. 7. Double-logarithmic plot of the variation in the rate at which VR(P, 400 C, t)
increases with time versus VR at 400 C and 2, 10 and 20 kbar. The lines are
calculated using the equation ln (dVR(P, 400 C, t)/dt) = ln (n(P, 400 C) k(P,400 C))
+ (1  n(P, 400 C)1) ln VR(P, 400 C, t) (Eq. (5)) and the best estimates for the
constants n(P, 400 C) and k(P, 400 C) given in Table 4. ln VR(P, 400 C, t) is given by
Eq. (3). The three lines intersect at VR = 1.44% and dVR(P, 400 C, t)/
dt = 7.00  106 % s1.

R. Le Bayon et al. / Organic Geochemistry 42 (2011) 340355

maturation of vitrinite when VR < 1.44% and promotes it when


VR > 1.44%. However, a VR of 1.44% is attained after a very short
heating time t of 12 h at 20 kbar, 7 h at 10 kbar and 4 h at
2 kbar and 400 C from a starting VR equal to 0%. Thus, the retardation of VR increase is insignicant for vitrinite maturation occurring in geological setting at 400 C. Consequently, pressure
promotes VR for geological heating time at 400 C, even though
pressure retards the initial VR evolution kinetics.
Several key features of the evolution of the constants n(P,
400 C) and k(P, 400 C) with pressure are evident from Fig. 6. In
particular, the three non-parallel lines of the ln VR(P, 400 C, t) evolution with ln t exhibit different positive slopes n(P, 400 C) that
are observed to increase with pressure having magnitudes ranging
between 0.07 and 0.21 (Table 4). The normalization constants n(P,
400 C) ln k(P, 400 C) decrease with increasing pressure. Nevertheless, the rate constants k(P, 400 C) increase with pressure (Table 4). The power law (Eq. (2)) and the linear (Eq. (3)) equations
with their associated constants n(P, 400 C) and k(P, 400 C) support all of the qualitative observations reported above and can be
observed in Fig. 6. In particular, these two equations accurately describe the expected increase in VR with time at each investigated
pressure because n(P, 400 C) > 0. Nevertheless, they call for a decrease in the rate of VR isobaric increase with time as n(P,
400 C) < 1. However, Eqs. (2) and (3) support that enhancing pressure results in a lesser deceleration of VR increase with time because n(P, 400 C) increases with pressure. Furthermore, the
higher pressure, the lower are the initial VR magnitude and evolution kinetics because the normalization constant n(P, 400 C) ln
k(P, 400 C) decreases with increasing pressure.
To describe the VR evolution rates at different temperatures and
pressures, Huang (1996), Dalla Torre et al. (1997) and Ernst and
Ferreiro Mhlmann (2004) favored an empirical rate equation in
the form of VR = ktn where t is time, k is a rate constant and n a constant unchanged over the experimental temperature and pressure
range. Even though this rate equation satisfactorily describes their
experimental data, we consider these studies to be limited because
they do not consider the activation time and experimental heat-up
effects and their associated corrections to the VR evolution rate
equation.
Results reported here emphasize the important role of n(P,
400 C) and k(P, 400 C) in the power law Eq. (2) and in its linear
transformation (Eq. (3)) to control the kinetic evolution of VR increase with time and pressure. Hence, we must next assess the
pressure dependence of n(P, 400 C) and k(P, 400 C).
4.2. Pressure dependence of the power law exponent n(P, 400 C)

0.30
400 C

0.25

n (P, 400 C)

350

0.20
0.15
0.10
0.05
r 2 = 0.999

0.00
0

10

15

20

25

P (kbar)
Fig. 8. Variation in power law exponent n(P, 400 C) with pressure at 400 C
(Table 4). The solid line is the least-squares best t to the data using the linear
equation n(P, 400 C) = n(0, 400 C) + BP (Eq. (6)). The error bars in n(P, 400 C) are
one standard deviation rn(P, 400 C). The error bars in pressure are 25 bar at 2 kbar
and 1 kbar at 10 and 20 kbar.

exponent n(P, 400 C) is successfully linearly related to P with a


correlation coefcient r2 > 0.999. We best estimate at 400 C the
two constants in Eq. (6) as n(0, 400 C) = 5.46 (0.43)  102 and
B = 7.55 (0.30)  106 bar1. The line of regression of the power
law exponent n(P, 400 C) on pressure P is plotted in Fig. 8.
The demonstration that n(P, 400 C) increases linearly with
pressure is of major importance to quantify the evolution of vitrinite reectance with time and pressure. It conrms all the pressure
effects observed above on the rate of VR increase with time by
means of n(P, 400 C). Furthermore, the recognition that n(P,
400 C) is a function of pressure begins to make apparent the pressure contribution to the rate Eq. (2).
4.3. Pressure dependence of the rate constant k(P, 400 C)
In addition to n(P, 400 C), the rate constant k(P, 400 C) seems
also to account for the pressure effect because k(P, 400 C) also varies with P. The k(P, 400 C) data are presented in Table 4 and are
plotted versus pressure P in Fig. 9a. The key feature revealed by Table 4 and by Fig. 9a is the huge increase in the rate constant k(P,
400 C) with pressure. We quantify the k(P, 400 C) variation as a
function of pressure P by best tting the k(P, 400 C) data over P
by an empirical power law equation in the form of

kP; 400  C CP D
Initially, the power law exponent was considered to be independent of pressure (Dalla Torre et al., 1997) and temperature (Huang,
1996; Dalla Torre et al., 1997; Ernst and Ferreiro Mhlmann, 2004).
In contrast, we have previously noted that the power law exponent
n(P, 400 C) increases with pressure. Therefore, it is of particular
interest to analyze the inuence of pressure on this variable.
Accordingly, we plot n(P, 400 C) (Table 4) versus pressure in
Fig. 8: A regular increase in the power law exponent with pressure
is evident. This graphically points out the linear pressure dependence of this exponent. We best t all the data in Fig. 8 to quantify
the variation of the power law exponent n(P, 400 C) as a function of
pressure P by least squares to a linear relation expressed as

nP; 400  C n0; 400  C BP

where the constant B with dimensions [bar]1 is the slope of the


regression line of n(P, 400 C) on P and the constant n(0, 400 C) is
the dimensionless power law exponent at which the pressure
would drop to zero at 400 C. The data are weighted assuming the
uncertainties in n(P, 400 C) given in Table 4. The power law

7
D

1

where C is a constant at 400 C with dimensions [bar] [time] and


D is the dimensionless exponent that characterizes the evolution of
k(P, 400 C) with P at 400 C. The data were weighted assuming the
uncertainties in k(P, 400 C) given in Table 4. The best estimates for
the two constants of Eq. (7) are C = 8.514 (108.89)  1089 barD s1
and D = 17.325 (1.245). Hence, the exponent D supports the observed increase in the rate constant k(P, 400 C) with pressure
(Fig. 9a) inasmuch as D > 0. The relationship between the variables
k(P, 400 C) and P is successfully described by the power law Eq. (7)
because this t shows a correlation coefcient r2 of 0.995. Despite this
excellent correlation coefcient, it is useful to test the t for the k(P,
400 C) data over pressure with a power law equation. For this purpose, we convert the nonlinear relation Eq. (7) to a linear one by taking its natural logarithm to give

ln kP; 400  C ln C D ln P

where D is the slope and ln C is the normalization constant of the


straight line ln k(P, 400 C) versus ln P. The straight line is calculated

R. Le Bayon et al. / Organic Geochemistry 42 (2011) 340355

VR = 0%. Kinetic models that do not consider effects of pressure,


effective heating time and heat-up to the nal temperature may
not accurately account for the evolution of vitrinite reectance in
natural systems. Our experiments require further investigations
at different temperatures in order to formulate a general equation
for vitrinite reectance evolution as a function of pressure, temperature and time. Hence, we regard our kinetic formulation as a step
toward this general equation. Such a PTt general equation would
lead to different pathways to calculate VR in agreement with measured values. Thus, VR cannot be converted directly into paleo-P
Tt conditions. Nevertheless, the number of possible solutions is
greatly diminished if information is available on temperature
(e.g., microthermometry based on uid inclusions, geothermal
investigations) and/or pressure (barometry on adjacent calc-silicate rocks) attained and/or the heating time maintained during
maturation (e.g., burial history in sedimentary basins, dating of
metamorphic minerals). This opens the possibility to use VR as a
tool to estimate the PTt history of vitrinite bearing terranes by
iteration based modeling. Calibration of the PTt history of vitrinite bearing terranes will be a great help to gain insight into their
geodynamic evolution and to improve hydrocarbon generation
modeling. Such tasks will be the subject of forthcoming
communications.

k (P, 400 C) 10 14 (s 1 )

400 C

r 2 = 0.995

0
0

10

15

20

25

P (kbar)
P (kbar)
2

20

30

20

400 C

40

10

14

10

19

10

30

50
60
70

k (P, 400 C) (s 1 )

lnk (P, 400 C)

10

4.4.1. How to use the power law formalism to model VR


It is of interest to make explicit how the power law formalism
has to be used to model VR for complex starting VR and PTt histories. This aims to avoid any misunderstanding and misuse of
such an equation. For this purpose, we postulate that the general
VR evolution rate equation obeys

r 2 = 0.995

80

VRP; T; t kP; TtnP;T


7

Fig. 9. (a) Variation in the rate constant k(P, 400 C) as a function of pressure at
400 C (Table 4). The solid curve is least-squares best t to the data using the power
law equation k(P, 400 C) = C PD (Eq. (7)). The error bars in k(P, 400 C) are one
standard deviation rk(P, 400 C). The error bars in pressure are 25 bar at 2 kbar and
1 kbar at 10 and 20 kbar. (b) Double-logarithmic plot of the rate constant k(P,
400 C) versus P at 400 C. The solid line is the calculated linear equation ln k(P,
400 C) = ln C + D ln P (Eq. (8)).

with Eq. (8), using the best estimates for the constants C and D given
above. A plot of this calculated straight line and of the negative retrieved ln k(P, 400 C) data versus ln P is shown in Fig. 9b. This plot
displays a good t of the data to the calculated straight line (Eq. (8))
inasmuch as the pairs (ln k(P, 400 C), ln P) lie on or close to this
straight line. This graphic check validates the linear relationship between ln k(P, 400 C) and ln P. Therefore, the use of a power law
equation (Eq. (7)) is conrmed to describe the increase in the rate
constant k(P, 400 C) with P at 400 C. The pressure sensitivity demonstration of the rate constant k(P, 400 C) is of interest because it
further makes apparent the pressure contribution to the VR evolution rate Eq. (2).
4.4. The pressure sensitive VR evolution rate equation discussion
Having dened the pressure dependence of the power law exponent n(P, 400 C) and the rate constant k(P, 400 C), we are now in a
position to make explicit the pressure effect on the VR evolution
kinetics. In this regard, we insert Eqs. (6) and (7) in Eq. (2) to get
 CBP

10

10

lnP (P in bar)

VRP; 400  C;t CPD tn0;400

351

We now have an expression for the variation of VR with time


and pressure at 400 C for Type III organic matter with initial

where the rate constant k(P, T) and the power law exponent n(P, T)
are both pressure and temperature dependent. This stems from the
demonstration that VR evolution is characterized by a power law
rate equation dependent on pressure (this study) and temperature
(Huang, 1996; Dalla Torre et al., 1997 and Ernst and Ferreiro Mhlmann, 2004). Furthermore, the kinetic parameters associated to this
power law formalism are pressure (this study) and temperature
dependent (Huang, 1996; Dalla Torre et al., 1997 and Ernst and
Ferreiro Mhlmann, 2004). The power law equation quantifying
VR evolution (e.g., our isothermal VR evolution rate Eq. (9)) is retrieved and formulated for a starting VR equal to 0% at t = 0. Consequently, the power law formalism with its associated rate constant
k and power law exponent n must be used to calculate VR only for a
starting VR equal to 0%. In nature, vitrinite precursors have reectance ranging between 0.1 and 0.2%. In addition, the starting
VR can display higher values resulting from one or several maturation events. However, the power law formalism (Eq. (10)) permits
the evaluation of VR attained after any heating time theating at PT
conditions (i.e., P2T2) different from the previous ones (i.e.,
P1T1) (Fig. 10) from any starting VR once a correction for heating
time is performed. For such VR estimation, we have to proceed in
two steps. First, we dene the heating time tstarting necessary at
P2T2 to increase VR from 0% to any starting VR previously reached
at P1T1 (i.e., VR(P1, T1)) (Fig. 10) with the power law Eq. (10) rearranged as

tstarting VRP1 ; T 1 nP2 ;T 2 =kP2 ; T 2

11

Second, VR attained after a heating time theating at P2T2


(i.e., VR(P2, T2)) from any starting VR previously reached at P1T1
(i.e., VR(P1, T1)) is dened by the power law formalism (Eq. (10))
as VR attained after a corrected heating time equal to
tstarting + theating from a starting VR equal to 0% at P2T2 (Fig. 10). This
is formulated as

352

R. Le Bayon et al. / Organic Geochemistry 42 (2011) 340355

400 C, even though we recognize a deceleration of the increase


in VR with time. For example, we calculate a VR equal to 6.82% at
400 C using the equation of Barker and Pawlewicz (1986) (ln
VR = 0.0078 Tmax(C) 1.2). To attain such a VR of 6.82 0.2% with
our VR rate Eq. (9), maturation at 400 C requires a heating time t of
613 Ma at 2 kbar, 475(100) years at 10 kbar and 6(1) years
at 20 kbar. Maturation with heating times other than those calculated using rate Eq. (9) displays different VR values from those calculated using the equation of Barker and Pawlewicz (1986). This
shows the limitations of the Barker and Pawlewicz VR equation

t starting + t heating
t starting

VR

t heating

P2 T2

3.0

t
: VR t evolution at conditions P2 T2
0 :0starting VR equal to 0%
1

: any starting VR obtained at P1 T1

2 : VR attained after any heating time t heating


at P2 T2 from any starting VR 1
Fig. 10. Sketch to illustrate how the power law formalism has to be used to
calculate VR for any starting VR and complex PTt history (see text).

VR (2 kbar,400 C, t )

(%)

a
2.5
2.0
1.5
1.0

2 kbar

0.5

: EASY%Ro

0.0
0

10

20

(%)

VR (10 kbar, 400 C, t )

4.4.2. Our experimental VR data and empirical rate equation:


Advantages, limitations and comparisons with other equations
Application of our empirical equation to Type I or II organic
matter is hazardous because these materials have different compositions. Indeed, Types I and II organic matter have high hydrogen
contents that inhibit VR enhancement.
Our VR rate equation is empirical and based on VR measured
after experimental maturation of vitrinite at 400 C under various
pressure and heating time conditions. Our rate equation is not
based on chemical kinetics because the chemical reactions in organic matter/vitrinite maturation are complex. Nevertheless, our
observations on VR evolution and our empirical VR kinetic equation
are only valid at 400 C. Further experimental investigations at different temperatures are necessary in order to understand the effects of pressure and time on VR at various temperatures and to
formulate a general equation for VR evolution as a function of pressure, temperature and time. Such a general equation will be a useful tool to model the VRTt conditions in vitrinite bearing
sedimentary basins and to estimate the PTt conditions in vitrinite bearing metamorphic terranes occurring in various tectonic
settings. For this purpose, our power law formalism has the advantage to be a simple and fast algorithm to model VRPTt evolution
in complex geological systems. This is of special interest in large
scale basin simulations that have to be performed and repeated
many times.
The earlier VR equations have several limitations. Barker (1983)
and Barker and Pawlewicz (1986) developed a VR equation based
on a statistical relationship between measured VR and temperature
proles in sedimentary basins. They argued that the geological
temperature regime is sufcient to obtain a stable VR. Hence, they
ignored the effect of heating time. However, we demonstrate in
this study that heating time is crucial in VR enhancement at

40

50

60

3.0

2.5
2.0
1.5
1.0

10 kbar

0.5

: EASY%Ro

0.0
0

10

15

20

25

30

t (days)
4.0

(%)

where tstarting is given by Eq. (11).


This demonstrates the potential of application of the power law
formalism to model VR from any starting VR and for complex PTt
histories. Furthermore, this brings out that the power law formalism is a simple and fast algorithm to calculate VR.

30

t (days)

12

VR (20 kbar,400 C, t )


nP2 ;T 2
VRP2 ; T 2 kP2 ; T 2 t heating t starting

3.5
3.0
2.5
2.0
1.5
1.0

20 kbar

0.5

: EASY%Ro

0.0
0

10

15

20

25

30

t (days)
Fig. 11. Comparison of VR modeled with EASY%Ro to VR obtained with experiments
(Table 3) as a function of effective heating time t at 400 C and (a) 2 kbar, (b) 10 kbar
and (c) 20 kbar. The gray thick curves are VR evolution with time modeled with
EASY%Ro at 400 C. The light curves are the least-squares best ts to the data using
the power law Eq. (2) that are displayed in Fig. 5. The error bars are one standard
deviation rVR(P, 400 C, t).

R. Le Bayon et al. / Organic Geochemistry 42 (2011) 340355

at 400 C despite the VR plateau calculated with our rate Eq. (9) for
geological heating times (613) at 2 kbar and 400 C that approximates the VR calculated with the equation of Barker and Pawlewicz
(1986). Nevertheless, Ernst and Ferreiro Mhlmann (2004) show
that the VR increase at 2 kbar and 200 C is stabilized rapidly with
heating time. Their long duration laboratory results demonstrate
that at constant PT conditions, a quasi-threshold VR value is
attained in a geologically insignicant time and thereafter VR
increases progressively more slowly with continued heating. Thus,
the equation of Barker and Pawlewicz (1986) might be relevant at
very low temperature conditions where the effects of time and
pressure should be less pronounced on the VR evolution rates.
Larter (1989) provided a VR equation based on the concentration of phenols released from vitrinite at different maturities during pyrolysis. A serious limitation of this approach is that the
amounts of phenols released from organic matter are easily quantied only in the range 0.45% < VR < 1.6%. Another equation is the
EASY%Ro method (Burnham and Sweeney, 1989; Sweeney and
Burnham, 1990), based on the rate of cracking of the most important volatile products from vitrinite. The generation of these compounds is described by a rst order Arrhenius equation and the
procedure follows the basic work of Karweil (1956) and improvements by Tissot and Espitali (1975). The use of both of the chemistry based VR equations is limited for high pressure conditions
because they do not consider the important pressure effect on VR
demonstrated in this study. It is of interest to compare the well
known kinetic model of VR evolution EASY%Ro with the experimental VR obtained in this study at 400 C. We calculate VR evolution
with time at 400 C with EASY%Ro and plot it in the experimental
VR(P, 400 C, t) versus t diagrams (Fig. 11). The rate model EASY%Ro
of Sweeney and Burnham (1990) provides VR that are higher than
the 2 kbar experimental data and lower than the 20 kbar runs. The
disparities between the model and the laboratory maturation are
less at 10 kbar: VR calculated with EASY%Ro are slightly lower than
the experimental VR. It is obvious that the upper temperature limit
is reached at 400 C for the use of EASY%Ro. This argues for a strict
use of EASY%Ro for temperature T  400 C. In addition, the disparities between modeled and experimental maturation change with
increasing pressure. It shows the importance to include pressure
as a parameter controlling the maturation in kinetic model of VR
evolution. We believe that our kinetic experiments on vitrinite
maturation may aid in model (e.g., EASY%Ro) improvement. However, the phenol kinetic and the popular EASY%Ro methods are useful tools to calibrate and model VR at very low temperature and
when the pressure effects on VR evolution rates can be neglected,
i.e., at very low pressure conditions.
This discussion points out several limitations in the already
existing VR equations and demonstrates the strength of experimentally derived empirical VR rate equation to model VR evolution.

5. Summary
The experimental results demonstrate that VR increases with
heating time at 400 C and at each investigated pressure. Despite
rapid initial kinetics, a deceleration of VR increase with time takes
place at each pressure. A variation in pressure effects on VR increase
kinetics is found. When VR < 1.44%, increasing pressure decreases
the rate of VR increase and hence retards the initial VR enhancement. Consequently, the lower pressure, the higher are the initial
VR magnitude and evolution kinetics. Nevertheless, the retarding
effect of pressure on VR increase diminishes with enhancing VR.
However, a VR of 1.44% is rapidly attained (only a few hours) at
400 C. Therefore, the retardation of VR increase is insignicant
for geological maturation at 400 C. When VR > 1.44%, increasing
pressure reduces the deceleration of VR enhancement with time

353

and thus results in a larger VR increase with time at 400 C. Obviously, heating time and pressure promote VR increase and the maturation of Type III organic material at 400 C. In addition, Type III
organic matter maturation is shown to rapidly occur in a dry conned system and does not require added water. Finally, a strong effect of the experimental heat-up on VR is identied and quantied,
even on experiments of very short duration. This information is
important because the experimental heat-up effect must be corrected in kinetic analysis.
The evolution of vitrinite reectance with time and pressure
at 400 C from an initial VR value of 0% is accurately described
by our new power law rate equation VR(P, 400 C, t) =
(k(P, 400 C) t)n(P, 400 C), where the exponent n(P, 400 C) increases linearly with pressure and the rate constant k(P, 400 C)
that obeys a power law equation with pressure, increases with
pressure. This VR evolution rate equation with its associated power
law exponent n(P, 400 C) and rate constant k(P, 400 C) accounts
for our experimental observations. In particular, this equation describes the expected increase in VR with time inasmuch as n(P,
400 C) > 0. Nevertheless, this equation calls for a deceleration of
VR increase with time because n(P, 400 C) < 1. Furthermore, the
VR power law rate equation shows the observed pressure effects
on VR evolution kinetics. Increasing pressure results in the lesser
deceleration of VR enhancement with time because n(P, 400 C) increases with pressure. With increasing pressure, the power law
rate equation displays the initial VR increase retardation when
VR < 1.44% and the larger VR increase with time when VR > 1.44%.
We regard our kinetic formulation as a step toward a general
equation describing VR evolution as a function of time, pressure
and temperature for Type III organic matter. Such VR rate equation
is empirical and based on VR measured after experimental maturation of vitrinite under various pressure, temperature and heating
time conditions. Our rate equation is not based on chemical kinetics because the chemical reactions in organic matter/vitrinite maturation are complex. We show the potential of the power law
formalism to model VR from any starting VR and for complex metamorphic and heating time histories by making explicit how to use
such a kinetic equation. The power law formalism is a simple and
fast algorithm to simulate VRPTt evolution. This is an advantage
to model large-scale sedimentary basins and metamorphic terranes. Kinetic models that do not consider the effects of effective
heating time, pressure and experimental heat-up to reach the desired run temperature may not accurately account for the evolution of vitrinite reectance in natural systems. Despite these
advances, signicant insights into VR evolution kinetics may
emerge through future experimental studies at different temperatures or/and in wet (water added) conned systems.
Acknowledgements
This research was supported by the Deutsche Forschungsgemeinschaft (DFG) Grant BA 3527/1-1. We thank T. Kautz for help in
the high-pressure lab, V. Bullatov for technical discussions during
machining of the piston-cylinder assemblies, W. Pttmann for analyzing the chemistry of the xylite sample, M. Dolezych for supplying the original xylite and S. Weinbruch for a check of the
manuscript. The manuscript greatly beneted from detailed comments by A. Burnham, R. Hill and an anonymous reviewer. R. Littke
provided useful reviews on an earlier draft of this paper.
Associate EditorKen Peters
References
rkai, P., Ferreiro Mhlmann, R., Suchy, V., Balogh, K., Sykorov, I., Frey, M., 2002.
Possible effects of tectonic shear strain on phyllosilicates: a case study from the

354

R. Le Bayon et al. / Organic Geochemistry 42 (2011) 340355

Kandersteg area, Helvetic domain, Central Alps, Switzerland. Schweizerische


Mineralogische und Petrographische Mitteilungen 82, 273290.
Armstrong, P.A., Chapman, D.S., Funnell, R.H., Allis, R.G., Kamp, P.J.J., 1996. Thermal
modelling and hydrocarbon generation in an active-margin basin: Taranaki
Basin, New Zealand. American Association of Petroleum Geologists Bulletin 80,
12161241.
Barker, C.E., 1983. Inuence of time on metamorphism of sedimentary organic
matter in selected geothermal systems, western North America. Geology 11,
384388.
Barker, C.E., 1989. Temperature and time in the thermal maturation of sedimentary
organic matter. In: Naeser, N.D., McCulloh, T.H. (Eds.), Thermal History of
Sedimentary Basins. Springer, New York, pp. 7398.
Barker, C.E., 1991. Implications for organic maturation studies of evidence for a
geologically rapid increase and stabilization of vitrinite reectance at peak
temperature: Cerro Prieto geothermal system, Mexico. American Association of
Petroleum Geologists Bulletin 75, 18521863.
Barker, C.E., Pawlewicz, M.J., 1986. The correlation of vitrinite reectance with
maximum temperature in humic organic matter. In: Butebarth, G., Stegena, L.
(Eds.), Paleogeothermics. Springer-Verlag, New York, pp. 7993.
Bonijoly, M., Oberlin, M., Oberlin, A., 1982. A possible mechanism for natural
graphite formation. International Journal of Coal Geology 1, 283312.
Bostick, N.H., 1974. Phytoclasts as indicators of thermal metamorphism, Franciscan
assemblage and Great Valley sequence (Upper Mesozoic), California. Geological
Society of America Special Papers 153, 117.
Burnham, A.K., Sweeney, J.J., 1989. A chemical kinetic model of vitrinite maturation
and reectance. Geochimica et Cosmochimica Acta 53, 26492657.
Carr, A.D., 2000. Suppression and retardation of vitrinite reectance, Part 1.
Formation and signicance for hydrocarbon generation. Journal of Petroleum
Geology 23, 313343.
Chandra, D., 1965. Reectance of coals carbonized under pressure. Economic
Geology 60, 621629.
Dalla Torre, M., de Capitani, C., Frey, M., Underwood, M.B., 1994. Very low-grade
metamorphism of shales from the Diablo Range, Franciscan Complex, California.
EOS 75, 699.
Dalla Torre, M., de Capitani, C., Frey, M., Underwood, M.B., Mullis, J., Cox, R., 1996.
Very low-grade metamorphism of shales from the Diablo Range, Franciscan
Complex, California, USA: new constraints on the exhumation path. Geological
Society of America Bulletin 108, 578601.
Dalla Torre, M., Ferreiro Mhlmann, R., Ernst, W.G., 1997. Experimental study on the
pressure dependence of vitrinite maturation. Geochimica et Cosmochimica Acta
61, 29212928.
Diessel, C.F.K., Brothers, R.N., Black, P.M., 1978. Coalication and graphitization in
high-pressure schists in New Caledonia. Contributions to Mineralogy and
Petrology 68, 6378.
Durand, B., Espitali, J., 1976. Geochemical studies on the organic matter from the
Douala Basin (Cameroon) II. Evolution of kerogen. Geochimica et
Cosmochimica Acta 40, 801808.
Ernst, W.G., Ferreiro Mhlmann, R., 2004. Vitrinite alteration rate as a function of
temperature, time, starting material, aqueous uid pressure and oxygen
fugacity laboratory corroboration of prior work. In: Hill, R.J., Leventhal, J.,
Aizenshtat, Z., Baedecker, M.J., Claypool, G., Eganhouse, R., Goldhaber, M., Peters,
K. (Eds.), Geochemical Investigations in Earth and Space Science: A Tribute to
Isaac R. Kaplan, No. 9. The Geochemical Society, pp. 341357.
Ferreiro Mhlmann, R., 2001. Correlation of very low grade data to calibrate a
thermal maturity model in a nappe tectonic setting, a case study from the Alps.
Tectonophysics 334, 133.
Frey, M., Teichmller, M., Teichmller, R., Mullis, J., Knzi, B., Breitschmid, A.,
Gruner, U., Schwizer, B., 1980. Very low-grade metamorphism in external parts
of the Central Alps: illite crystallinity, coal rank and uid inclusion data. Eclogae
Geologicae Helvetiae 73, 173203.
Frings, K., Lutz, R., de Wall, H., Warr, L.N., 2004. Coalication history of the Stefanian
Cinera-Matallana pull-apart basin, NW Spain: combining anisotropy of vitrinite
reectance and thermal modelling. International Journal of Earth Sciences 93,
92106.
Goodarzi, F., 1985. Optical properties of vitrinite carbonized at different pressures.
Fuel 64, 156162.
Hao, F., Yongchuan, S., Sitian, L., Qiming, Z., 1995. Overpressure retardation of
organic-matter maturation and petroleum generation: a case study from the
Yinggehai and Qiongdongnan basins, South China Sea. American Association of
Petroleum Geologists Bulletin 79, 551562.
Hao, F., Zou, H., Gong, Z., Yang, S., Zeng, Z., 2007. Hierarchies of overpressure
retardation of organic matter maturation: case studies from petroleum basins
in China. American Association of Petroleum Geologists Bulletin 91, 14671498.
Hill, R.J., Jenden, P.D., Tang, Y.C., Teerman, S.C., Kaplan, I.R., 1994. Inuence of
pressure on pyrolysis of coal. In: Mukhopadhyay, P.K., Dow, W.G. (Eds.),
Vitrinite Reectance as a Maturity Parameter: Applications and Limitations.
American Chemical Society Symposium Series, vol. 570. American Chemical
Society, Washington, DC, pp. 161193.
Hood, A., Gutjahr, C.M., Heacock, R.L., 1975. Organic metamorphism and the
generation of petroleum. American Association of Petroleum Geologists Bulletin
59, 986996.
Horseld, B., Disko, U., Leistner, F., 1989. The microscale simulation of maturation:
outline of a new technique and its potential applications. Geologische
Rundschau 78, 361374.
Hryckowian, E., Dutcher, R.R., Dachille, F., 1967. Experimental studies of anthracite
coals at high pressures and temperatures. Economic Geology 62, 517539.

Huang, W.L., 1996. Experimental study of vitrinite maturation: effects of


temperature, time, pressure, water, and hydrogen index. Organic Geochemistry
24, 233241.
Hunt, J.M., 1979. Petroleum Geochemistry and Geology. Freeman, San Francisco.
Hunt, J.M., 1996. Petroleum Geochemistry and Geology, second ed. Freeman and
Co., San Francisco.
Hunt, J.M., Lewan, M.D., Hennet, R.J.C., 1991. Modeling oil generation with time
temperature index graphs based on the Arrhenius equation. American
Association of Petroleum Geologists Bulletin 795, 807.
Hutton, A.C., Cook, A.C., 1980. Inuence of alginite on the reectance of vitrinite
from Joadja, NSW, and some other coals and oil shales containing alginite. Fuel
59, 711714.
Johnsson, M.J., Howell, D.G., Bird, K.J., 1993. Thermal maturity patterns in Alaska:
implications for tectonic evolution and hydrocarbon potential. American
Association of Petroleum Geologists Bulletin 77, 18741903.
Karweil, J., 1956. Die Metamorphose der Kohlen vom Standpunkt der Physikalischen
Chemie. Zeitschrift fr der Deutschen Geologischen Gesellschaft 107, 132139.
Kerrick, D.M., 1987. Cold-seal systems. In: Ulmer, G.C., Barnes, H.L. (Eds.),
Hydrothermal Experimental Techniques. Wiley, New York, pp. 293323.
Kisch, H.J., 1987. Correlation between indicators of very low-grade metamorphism.
In: Frey, M. (Ed.), Low Temperature Metamorphism. Blackie, New York, pp. 226
300.
Landais, P., Monin, J.C., Monthioux, M., Poty, B., Zaugg, P., 1989. Dtermination
exprimentale de lvolution dindicateurs de maturation de la matire
organique. Application aux krognes du Toarcien du Bassin de Paris.
Comptes Rendus de lAcadmie des Sciences (Paris) 308, 11611166.
Landais, P., Michels, R., Elie, M., 1994. Are time and temperature the only constraints
to the simulation of organic matter maturation? Organic Geochemistry 22,
617630.
Larter, S., 1989. Some pragmatic perspectives in source rock geochemistry. Marine
and Petroleum Geology 5, 194204.
Law, B.E., Nuccio, V.F., Barker, C.E., 1989. Kinky vitrinite reectance well proles:
evidence of paleopore pressure in low-permeability, gas-bearing sequences in
rocky mountain foreland basins. American Association of Petroleum Geologists
Bulletin 73, 9991010.
Le Bayon, R., Brey, G.P., Ernst, W.G., Ferreiro Mhlmann, R., 2007. Experimental
kinetic study of organic carbonaceous material maturation: an appraisal of
pressure and time effects on vitrinite reectance at 400 C. Geochimica et
Cosmochimica Acta 71, A552 (Abstract. Goldschmidt Conference, Cologne).
Lenz, O.K., Riegel, W., 2001. Isopollen maps as a tool for the reconstruction of a
coastal swamp from the Middle Eocene at Helmstedt (northern Germany).
Facies 45, 177194.
Lewan, M.D., 1993. Laboratory simulation of petroleum formation: hydrous
pyrolysis. In: Engel, M.H., Macko, S.A. (Eds.), Organic Geochemistry. Plenum
Publishing, New York, pp. 419442.
Lewan, M.D., Winters, J.C., McDonald, J.H., 1979. Generation of oil-like pyrolysates
from organic rich shales. Science 203, 897899.
Lopatin, N.V., 1971. Temperature and geologic time as factors in coalication.
Izvestiya Akademii Nauk SSSR Seriya Geologicheskaya 3, 95106 (in Russian).
Mastalerz, M., Wilks, K.R., Bustin, R.M., Voss, J.V., 1993. The effect of temperature,
pressure and strain on carbonization in high-volatile bituminous and
anthracitic coals. Organic Geochemistry 20, 315325.
McTavish, R.A., 1978. Pressure retardation of vitrinite diagenesis, offshore northwest Europe. Nature 171, 648650.
Monthioux, M., 1988. Expected mechanisms in nature and in conned-system
pyrolysis. Fuel 67, 843848.
Monthioux, M., Landais, P., Monin, J.C., 1985. Comparison between natural and
articial maturation series of humic coals from the Mahakam delta, Indonesia.
Organic Geochemistry 8, 275292.
Murchison, D.G., Cook, A.C., Raymond, A.C., 1985. Optical properties of organic
matter in relation to thermal gradients and structural deformation.
Philosophical Transactions of the Royal Society 315, 157186.
Parris, T.M., Burrus, R.C., OSullivan, P.B., 2003. Deformation and the timing of gas
generation and migration in the eastern Brooks Range foothills, Arctic National
Wildlife Refuge, Alaska. American Association of Petroleum Geologists Bulletin
87, 18231846.
Petersen, H.I., Rosenberg, P., 1998. Reectance (suppression) and source rock
properties related to hydrogen-enriched vitrinite in Middle Jurassic coals,
Danish North Sea. Journal of Petroleum Geology 21, 247263.
Price, L.C., 1983. Geologic time as a parameter in organic metamorphism and
vitrinite reectance as an absolute paleogeothermometer. Journal of Petroleum
Geology 6, 538.
Price, L.C., Barker, C.E., 1985. Suppression of vitrinite reectance in amorphous rich
kerogen a major unrecognized problem. Journal of Petroleum Geology 8, 59
84.
Raymond, A.C., Murchison, D.G., 1991. Inuence of exinitic macerals on the
reectance of vitrinite in carboniferous sediments of the Midland Valley of
Scotland. Fuel 70, 155161.
Ritter, U., 1984. The inuence of time and temperature on vitrinite reectance.
Organic Geochemistry 6, 473480.
Robert, P., 1980. The optical evolution of kerogen and geothermal histories applied
to oil and gas exploration. In: Durand, B. (Ed.), Kerogen, vol. 12. Technip, Paris,
pp. 385414.
Roksandic, M.M., 1986. Dynamic interpretation of organic matter maturation and
evolution of oil-generative window: discussion. American Association of
Petroleum Geologists Bulletin 70, 10081010.

R. Le Bayon et al. / Organic Geochemistry 42 (2011) 340355


Stach, F., Mackowsky, M.-T., Teichmller, M., Taylor, G.H., Chandra, D., Teichmller,
R., 1982. Textbook of Coal Petrology. Gebrder Borntraeger, Stuttgart.
Suzuki, N., Matsubayashi, H., Waples, D.W., 1993. A simpler kinetic model of
vitrinite reectance. American Association of Petroleum Geologists Bulletin 77,
15021508.
Sweeney, J.J., Burnham, A.K., 1990. Evaluation of a simple model of vitrinite
reectance based on chemical kinetics. American Association of Petroleum
Geologists Bulletin 74, 15591570.
Taylor, G.H., Teichmller, M., Davies, A., Diessel, C.F.K., Littke, R., Robert, P., 1998.
Organic Petrology. Gebrder Borntraeger, Berlin.
Teichmller, M., Teichmller, R., 1954. Die stofiche und strukturelle
Metamorphose der Kohle. Geologische Rundschau 42, 265296.
Teichmller, M., Teichmller, R., 1979. Diagenesis of coal (coalication). In: Larsen,
G., Chilingar, G.V. (Eds.), Diagenesis in Sediments and Sedimentary Rocks.
Elsevier, Amsterdam, pp. 207246.
Tissot, B., Espitali, J., 1975. Lvolution thermique de la matire organique des
sdiments: applications dune simulation mathmatique. Revue de lInstitut
Franais du Ptrole 30, 743777.
Tissot, B.P., Welte, D.H., 1984. Petroleum Formation and Occurrence. SpringerVerlag, New York.
Tissot, B.P., Pellet, R., Ungerer, Ph., 1987. Thermal history of sedimentary basins,
maturation indices, and kinetics of oil and gas generation. American Association
of Petroleum Geologists Bulletin 71, 14451466.

355

Underwood, M.B., Shelton, K.L., McLaughlin, R.J., Laughland, M.M., Solomon, R.M.,
1999. Middle Miocene paleotemperature anomalies within the Franciscan
Complex of northern California: thermo-tectonic responses near the
Mendocino triple junction. Geological Society of America Bulletin 111, 1448
1467.
van Krevelen, D.W., 1953. Physikalische Eigenschaften und chemische Struktur der
Steinkohle. Brennstoff-Chemie 33, 167182.
Vandenbroucke, M., Behar, F., San Torcuato, A., Rullkter, J., 1993. Kerogen
maturation in a reference kerogen type II series: the Toarcian shales of the
Hills Syncline, NW Germany. Organic Geochemistry 20, 961972.
Waples, D.W., 1980. Time and temperature in petroleum formation: application of
Lopatins method to petroleum exploration. American Association of Petroleum
Geologists Bulletin 64, 916926.
Waples, D.W., Kamata, H., Suizu, M., 1992. The art of maturity modelling. Part 1:
nding a satisfactory geologic model. American Association of Petroleum
Geologists Bulletin 76, 3148.
Wenger, L.M., Baker, D.R., 1987. Variations in vitrinite reectance with organic
facies examples from Pennsylvanian cyclothems of the mid-continent, USA.
Organic Geochemistry 11, 411416.
Wolf,
M.,
1978.
Inkohlungsuntersuchungen
im
Hunsrck,
Rheinishe
Schiefergebirge. Zeitschrift fr der Deutschen Geologischen Gesellschaft 129,
217227.

You might also like